Вы находитесь на странице: 1из 20

Graduate Student Association Research Fair

University of Nevada, Reno


Spring 2006

2D AND 3D NUMERICAL SIMULATIONS OF COMBINED NATURAL


CONVECTION AND RADIATION HEAT TRANSFER BETWEEN A
CIRCULAR CYLINDER AND OUTER SQUARE ENCLOSURE
Pablo E. Araya Gómez
University of Nevada, Reno
Mechanical Engineering Department
Email: arayap@unr.edu

ABSTRACT

Two-dimensional and three-dimensional computational models of a horizontal cylinder in

a square enclosure are constructed based on geometric and material characteristics of a

horizontal spent nuclear fuel rod under storage and transport conditions. The rod generates heat

which is transported outwards through the backfill gas environment by conduction, convection

and radiation. Heat transfer simulations are performed to predict temperatures and velocities

within the domain for varying heat generation rates with a uniform enclosure wall temperature of

25°C, and with nitrogen backfill gas. Different sets of simulations model conduction/radiation

and natural convection/radiation transport across the gas filled region to assess the importance of

the different transport phenomena. Results show that simplified 2D models can be used with

little error in results and that the 2D simulations tend to over predict resulting temperatures by as

much as 5% for the higher heat loads when compared to 3D models. Differences between model

with and without radiation heat transfer show that it has significant effects in reducing

temperatures by as much as 37%.

INTRODUCTION

The steady state heat transfer of buoyancy induced flows between objects and their

enclosures has become of increasing interest and focus by researchers. This because many new

1
applications and technologies have become more predominant, such as solar collectors and

receivers, insulation and flooding protection for buried pipes, electronic packaging and air-

conditioning applications. However, the main focus of this paper is tied to the study of the

cooling process of spent nuclear fuel (SNF) rods that are stored within a square basket, backfilled

with a non-oxidizing gas. This problem primarily consists of cylindrical heated objects

contained within a square enclosure in a horizontal orientation. A spent nuclear fuel assembly is

composed of clad uranium dioxide rods arranged in a square horizontal array. They are enclosed

by a square basket structure. The rods generate heat which is transported towards the basket

walls by conduction, natural convection and radiation. Since radiation heat transfer is of primary

importance in spent nuclear fuel storage, it is included in this study.

The study of natural convection in enclosures has been conducted by many researchers

focused primarily on concentric and eccentric annuli [Shu et. al, 2001, Ding et. al, 2005.] Most

early studies are experimental [Bishop et. al, 1968] due to a lack of computational power to solve

the complexities present in buoyancy induced flows. Some authors have focused on varying

shapes of the inner heated object in several backfill gases to determine general relationships

between the inner rod surface temperatures and Nusselt numbers with respect to flow Rayleigh

numbers [Warrington and Powe, 1984]. Lieberman and Gebhart [1969] went further by

experimentally studying horizontal arrays of heated elements whereas Tokura et. al [1983]

performed experiments on vertical arrays of heated elements. Early numerical works by Prussa

and Yao [1983] and Vasseur et. al [1983] were focused primarily on the positioning of the inner

cylinder in a cylindrical enclosure. Ghaddar [1992] presents the earliest numerical work found

in the reviewed literature where a rectangular enclosure is explicitly considered. Ghaddar

studied the natural convection resulting from a very small heated cylindrical object in a

rectangular enclosure to construct a functional relationship between the Nusselt numbers and

2
Rayleigh Numbers. Ekundayo et. al [1998] studied experimentally the position of a small

cylindrical heated object in a square enclosure. Moukalled and Acharya [1996] also considered a

square enclosure and obtained Nusselt numbers as a function of Rayleigh numbers for various

cylinder radii. Liu et. al [1996] determined that the buoyancy induced flow from a heated

cylinder within a cylindrical and rectangular enclosure is primarily dependent on the Rayleigh

number.

Sasaguchi et. al [1998] conducted numerical simulation of the transient cooling of a

heated cylinder immersed in a square enclosure filled with water. They are among the few

authors to consider a transient process in a numerical study. Most other authors consider steady

state processes because most applications involve objects that continually produce a small and

constant heat load at levels that cannot result in unstable buoyancy effects. Also, authors mostly

consider the flow regime as steady because the Rayleigh numbers obtained from their

applications are far below the values in which transient instabilities appear from buoyancy

induced flows. It is unclear at which heat loads, the instabilities arise. Gebhart et. al [1988]

refer to a value of flux Grashof number as low as 380 in nitrogen backfill as the point in which

instabilities begin to appear. However this is for buoyancy induced flows along an open vertical

heated wall. A better reference point for a heated cylinder in an enclosure is given by Bishop et.

al [1968] who determined experimentally that a Grashof number of GrD = 1.57x105, begins to

show effects of transient instability. We will see that, for our application, the maximum GrD is

1.01x104, and therefore, stability and transient effects may not be of concern. However we must

keep in mind that this study is on a geometry that differs with respect to that used by Bishop et.

al.

The authors in the reviewed literature do not consider the effects of radiation heat transfer

between the heated object and the outer enclosure wall. It is clear that experimental studies

3
include radiation as a heat transport phenomena, but the authors have not referred to it or

indicated the emissivities of the surfaces. More recent studies of spent nuclear fuel storage have

considered radiation in enclosures, but these fail to consider natural convection [Bahney and

Lotz, 1996.]

The objective of this study is to simulate 2D and 3D natural convection and radiation heat

transfer in a cylinder within a square enclosure in order to observe the differences between the

different models and to see how the presence of radiation affects the transport of energy in the

enclosure. Since this study is related to heat transfer in spent nuclear fuel storage, the cylindrical

uranium dioxide fuel rod with zircaloy cladding and enclosure are fully modeled and appropriate

material properties are used. The dimensions, temperatures and heat generation rates used are

also consistent with spent nuclear fuel assemblies. Nuclear fuel assemblies typically include an

array of cylindrical fuel rods. However, in order to benchmark our methods and study the

problem in its most basic form, we consider only one rod.

NUMERICAL METHODS

Figure 1 shows the two dimensional domain used in this study which consists of one

uranium fuel rod housed in a zircaloy channel. The rod is composed of a uranium dioxide pellet

of 12.37 mm in diameter. This pellet is covered by a 0.8128 mm thick zircaloy cladding, with a

0.00304 mm gap separation between the pellet and cladding. The outer enclosure is a square

zircaloy channel wall of 134 mm side length. The dimensions and material properties for this

model are based on a General Electric 7x7 Boiling Water Reactor SNF assembly [Bahney and

Lotz, 1996.] Zircaloy is set as the enclosure material in cases where the spent nuclear fuel is

shielded by a channel. The gap and backfill regions are filled with nitrogen.

Temperature, velocity and pressure results are obtained by solving the steady state

momentum and energy equations using a finite volume method for natural convection with and

4
without radiation heat transfer. The steady state conservation of momentum and energy

equations are solved with a second order upwind scheme, to obtain velocity, pressure and

temperature fields. All of the simulations are run for conduction within the rods with natural-

convection and with radiation/natural-convection to quantify the effects of radiation and

buoyancy on resulting temperatures and velocities.

The governing equations solved for the 2D and 3D models are the continuity (1),

momentum (2) and energy (3) equations.

∂[ρu ] ∂[ρv] ∂[ρw ]


+ + =0 (1)
∂x ∂y ∂z

∂[ρuu ] ∂[ρuv] ∂[ρuw ] ∂p ∂  2  ∂u ∂v ∂w  ∂   ∂u ∂v  ∂   ∂u ∂w 


+ + =− +  µ 2 − −  + µ +  + µ +  (2a)
∂x ∂y ∂z ∂x ∂x  3  ∂x ∂y ∂z  ∂y   ∂y ∂x  ∂z   ∂z ∂x 

∂[ρvu ] ∂[ρvv] ∂[ρvw ] ∂p ∂   ∂u ∂v  ∂  2  ∂v ∂u ∂w  ∂   ∂v ∂w 


+ + =− + µ +  +  µ 2 − −  + µ +  + ρg (2b)
∂x ∂y ∂z ∂y ∂x   ∂y ∂x  ∂y  3  ∂y ∂x ∂z  ∂z   ∂z ∂y 

∂[ρwu ] ∂[ρwv] ∂[ρww ] ∂p ∂   ∂u ∂w  ∂   ∂v ∂w  ∂  2  ∂w ∂u ∂v 


+ + =− + µ +  + µ +  +  µ 2 − −  (2c)
∂x ∂y ∂z ∂z ∂x   ∂z ∂x  ∂y   ∂z ∂y  ∂z  3  ∂z ∂x ∂y 

∂[u (ρh )] ∂[v(ρh )] ∂[w (ρh )] ∂  ∂T  ∂  ∂T  ∂  ∂T 


+ + = k + k + k (3)
∂x ∂y ∂z ∂x  ∂x  ∂y  ∂y  ∂z  ∂z 

In equation (3) h is the enthalpy which is related to temperature by dh = CpdT. The gravitational

force is imposed as g = -9.8 m/s2 in y. The heat dissipation and pressure work terms of the

energy equation (3) are considered negligible and therefore not included in the energy equation

[Yang and Lloyd, 1985]. The 2D models are solved by neglecting the terms in z and without

equation (2c). It is important to note that in the governing equations the density, specific heat

and thermal conductivity of the nitrogen gas backfill will be considered as functions of the

temperature. This is especially important for the density, since the buoyancy forces that result in

natural convection are a result of changes in density.

Radiation is solved for gray diffuse surfaces using the surface-2-surface (S2S) method.

5
This method is recommended for enclosures with non participating media and involves the

exchange of energy only between walls. The following radiation equation is solved for at each

surface, k.

N
J k = E k + ρk ∑F J
j=1
kj j (4)

In this expression Jk is the radiosity, ρk = 1-εk is the reflectivity of surface k, Ek = εkσTk4 the

emissive power and Fkj the view factor between surfaces k and j. Fkj is calculated by

1 cos θ k cos θ j
Fkj =
Ak ∫∫
Ak A j
πr 2
δ kj dA k dA j (5)

In this expression Ak is the area of surface k, r the distance between surfaces and θ the angle

between the normal and the vector from surface to surface for each respective surface k.

Equation (4) is a system of N x N equations, where N is the total number of surfaces. Note, that

in order for there to be a transfer of energy between surfaces, there has to be a temperature

difference and the surfaces must be able to “see” each other. This means that δkj = 1 if the

surfaces dAk and dAj are visible to each other; otherwise δkj=0. Since the exchange of radiation

depends on T4 it is apparent that the transfer of energy by radiation becomes more important than

other modes as the temperature difference between two surfaces increases.

The boundary conditions used on the 2D and 3D models are of uniform temperature Tw =

298.15 [K] on the outer enclosure and varying heat generation rates q from 4550 [W/m3] to

45500 [W/m3]. In the 3D model, the ends of the rod are considered adiabatic. The heat loads are

compatible with expected Uranium rod heat generation from spent nuclear fuel.

COMPUTATIONAL CODE

The governing equations are solved for using the commercial FLUENT 6.2 code which

uses a finite volume method to solve the discretized governing equations. The 2D and 3D grids

6
are constructed and solved for with double precision. The steady segregated solver is used since

the temporal terms are not considered and a second order upwind scheme is used for solving the

momentum and energy equations.

COMPUTATIONAL TEST CASE

We can see in Figure 1 one of the grids used for the simulations which consists of 20200

quad elements using a mapped and sub-mapped scheme where the near wall regions are meshed

using a growth factor of 1.04. This grid was selected by running several cases with varying grid

resolution. Since we do not have an analytical solution to compare to, we compare all solution to

that obtained with the finest grid composed of 36520 elements. All simulations were initialized

to zero velocity for u, v and w and initial temperature of 298.15 K. Figure 2 presents the grid

convergence for the maximum heat load of q = 45500 W/m3. It shows that, as the number of

elements is increased, the three variables that are monitored tend to converge. The three

variables selected are ∆Tc,max, maximum cladding to wall temperature difference, Smax, maximum

flow speed (velocity magnitude) and NuL, global Nusselt number. These are representative of

the continuity, momentum and energy equations. The variation of each of the variables, φ, with

respect to the reference grid D is examined for different resolutions. That is, for example, in case

of Nu, φ/φD = Nu/NuD. Grid B shows differences smaller than 0.4% in all monitored variables

with respect to grid D. This is considered acceptable, since grid B runs at ¼ of the

computational time as grid D and therefore grid B is chosen for further studies.

A simple 2D model (not shown) is also constructed using the same discretization scheme

as Grid B. The simple model replaces the rod components in the full 2D model, from Fig. 1, by

a simple cylinder composed of UO2 with no gap or cladding. The effects of the cladding are

imposed on the simple model as a wall resistance by adding a thickness of 0.8128 mm as a wall

boundary condition with zircaloy properties.

7
Figure 3 shows the 3D domain and grid selected to run simulations. The simple 2D

model has been used as a base for this 3D model. In the 3D model, the fuel rod has been

modeled without a gap and the cladding is only a wall property. The gap thickness is considered

small enough that any resistance it may cause will be considered negligible. Also the zircaloy

cladding is not meshed or modeled. It is assumed that if the conduction across the cladding is

principally radial, we can substitute it by a resistance that is input into the model as a boundary

conditions where Zircaloy is defined as the surface material with a thickness of 0.8128 mm. The

results will show whether or not these simplifications are valid or not. It is important to note that

stainless steel end plates are added to the model. End plates were not considered in the 2D

model. Also, a length of 610 mm is selected because it corresponds to the typical measure

between grid spacers in a spent nuclear fuel assembly [Lovett, 1991]. The model in Figure 3 has

its origin at the geometric center with x as the horizontal, y vertical and z axial coordinates. As

in the 2D case the 3D simulations were initialized at zero velocities and initial temperature of

25°C (298.15 K).

For the 3D model, three grids of different resolution are constructed using hex-mapped

meshing. To construct each 3D grid, a 2D face grid was created for the front end plate and then

extruded with a double sided growth factor of 1.04 in z. The same convergence criteria is used

for 3D as is 2D simulations.

Figure 4 shows temperature profiles for a maximum heat load of 55815 W/m3 which

corresponds to the same heat load as in the 2D grid convergence. The profiles are plotted along

a line placed on the axial center section plane of the geometry and oriented horizontally, 0.02 m

above center.

We can see that all of the profiles in Fig. 4 are very similar except at the center point.

However the maximum error in ∆T = T-Tw at the center is 5 % for the mesh with 103320

8
elements and considered acceptable. Therefore the mesh with 103320 elements is selected for

further studies.

RESULTS AND DISCUSSION

The 2D and 3D simulations are run to obtain velocity and temperature results. Figure 5

shows the temperature and velocity contours magnitude for the different models at the same heat

generation rate of Q = 5.2 Watt.

The simulations run for Figure 5 are of conduction and natural convection in the simple

2D model that does not include cladding and gap, the full 2D model with cladding and gap and

the 3D model. The similarity in temperature and velocity distributions between the different

models is evident. However it is important to note the maximum values of temperature predicted

by the 2D models are 0.24 % and 0.59 % higher for the 2D simple model and the 2D full model

respectively when compared to the 3D model. Also we can see that there is an attenuation of the

velocity contours for the 3D model. This is because a small portion of the momentum is

transferred in the axial direction causing an attenuation of the gradients. However, all maximum

velocity values are the same.

To visualize the axial effects in the 3D model we plot section planes with temperature

and velocity contours. Figure 6 shows a longitudinal bisecting plane and three sections of the

model with a heat load of Q = 5.2 W.

It is important to note that since the problem is symmetrical with respect to the xy plane

at z = 0, the solutions for z = -0.3 is the same as for z = 0.3. We can see how the temperature and

velocity contours vary for different sections along the z axis. Even though there are some effects

of the end plates on local velocities and temperatures, the center section is unaffected. We can

see axial effects on the plot of W, where underneath the rod there are some outward flows in z.

However, these are less than 10% of the total velocity magnitude, S.

9
The different 2D and 3D models can be compared further by varying the heat generation

in the rod. Figure 7 shows how the maximum cladding to enclosure wall temperature difference

is affected by a varying heat load.

We can see in Figure 7 that the 2D simple and 2D full models yield very similar results.

This means that it is feasible to replace the cladding and gap of the fuel rod with a simple wall

thermal resistance. When compared to the 3D models, the 2D models tend to over-predict

temperature to wall differences by as much as 5% for the highest heat load, Q= 5.2W. When

radiation heat transfer is added to the model, we can see that it has significant effects by lowering

temperatures by as much as 37% for the highest heat load Q = 5.2 W.

Figure 8 shows similar tendencies for the maximum velocity magnitude, Smax, as a

function of heat loads. In this case the difference between 2D and 3D models is hardly

noticeable. Again the presence of radiation significantly affects Smax by reducing it, because a

portion of the heat energy is transferred directly between the surfaces and does not affect the gas

backfill temperatures. Therefore buoyancy effects become less important.

To verify that the flows that have been studied are effectively steady and free of

instabilities and transient effect we can calculate the Grashof Number by GrD = gβ∆TD3/ν2 and

plot with respect to heat load. The rod diameter D is used as the reference length in order to

compare [Bishop et. al,1968].

The Grashof number is the ratio of the buoyancy force which moves the fluid with

respect to the viscous forces which retards it. We can see in Figure 9 that higher heat loads result

in higher GrD, as expected. Also, when radiation is imposed, a portion of the heat energy is no

longer transferred by natural convection and therefore the GrD are significantly lower.

The results from the present study are compared to previous experimental and numerical

work, by means of Nusselt and Rayleigh numbers. However, careful treatment of results must be

10
emphasized, since previous experimental works have considered a variety of geometries and

reference values. For example, the Nusselt is defined by NuL = hL/k. The values of heat transfer

coefficients, h, are obtained from the simulation results, and the conductivity, k, is specified as a

material property. L however, is a reference length which has been specified differently among

various authors who have considered it to be enclosure width, cylinder radius, cylinder to

enclosure gap, etc. Also, numerical works have considered different boundary conditions for the

rod. This study uses a volumetric heat generation rate for the solid rod whereas others have used

isothermal and wall heat flux on the rod surface.

Figure 10 shows a comparison of the 2D and 3D simulation results with previous

experimental results by Warrington and Powe (1994) and Numerical results by Ghaddar (1991.)

Both authors mentioned published correlated data functions of NuL vs RaL. It is important to

note that the reference length used to calculate the Nusselt numbers and Rayleigh Numbers is

based on the hypothetical gap between concentric spheres with equivalent volumes as the

enclosure and cylinder. Overall the results are in the same order of magnitudes and tendencies.

However, the results from this study show that NuL increases at a higher rate with RaL when

compared to both previous studies. This may be due to radiation effects or geometric differences

between our models and the previous authors. It is important to note that there are some

differences with respect to the model employed by the authors. Warrington and Powe (1994)

solved experimentally for object enclosed in a square box of equal lengths and widths, making

the 3D axial effects much more predominant. Also, their results consist of many data points for

four different geometries with average deviation of 14%. They solved for large cylinder radius

resulting in constricted buoyancy effects due to the small length to radius ratio, L/Ri. Ghaddar

(1991) obtained 2D numerical results for a very small cylinder of set from the center of a large

rectangular enclosure and modeled the heat load as a heat flux at the cylinder wall. More models

11
with varying cylinder diameters need to be studied in order to make better comparison to

previous work.

CONCLUSIONS

Natural convection and radiation simulations are conducted using commercial finite

volume software to study the heat transfer from a heat generating cylinder within a square

enclosure. The simulations consider aspects that have only been partially treated in previous

studies such as the relationship between combined natural convection and radiation heat transfer

and a comparison between 2D and 3D models. Two different 2D models are constructed to

verify that rod model simplifications can be applied without significant effects over results. Also

3D models are constructed and compared to 2D results. These show that there is little difference

when comparing 2D and 3D models. 2D models tend to over-predict temperatures by 5% for the

highest applied heat loads. Applying radiation heat transfer however, results in a maximum

decrease of 37% in temperature results when compared to cases that do not consider radiation.

Results are compared to those from previous studies. However, a more detailed study of

geometric parameters needs to be done in order to generate better comparisons with previous

publications.

Future studies will include numerical as well as experimental testing of these models.

These include a geometric study of varying lengths and diameters, the use of different backfill

gases such as Helium and Argon and the incidence of replacing the single cylinder with multiple

heating objects. Our challenge is to be able to relate the single cylinder problems to problems

that include, for example, square cylinder arrays of multiple heated rods that represent more

closely the storage of spent nuclear fuel assemblies.

12
NOMENCLATURE

GrL Grashof Number (gβ∆TL3/ν2)


NuL Nusselt number (hL/k)
L Rod Length
Pr Prandtl Number (0.71 for Nitrogen)
RaL Rayleigh Number (PrGrL)
ρ Density [kg/m3]
Cp Specific Heat [j/kgK]
k Thermal conductivity [W/mK]
µ Viscosity [kg/ms]
Tw Basket wall temperature [ºC]
Tc Maximum surface cladding temperature [ºC]
∆T Tc–Tw; Maximum cladding to wall temperature difference [ºC]
Q Total heat load or heat generation [W]
q Volumetric heat load [W/m3]
ε Emissivity
Ek Emissive power of surface k [W]
S Flow speed [cm/s]
g Gravitational force [-9.8 m/s2 in y]

REFERENCES

Shu, C., Xue, H. and Zhu, Y. D., 2001, “Numerical Study of Natural Convection in an Eccentric
Annulus Between a Square Outer Cylinder and a Circular Inner Cylinder Using DQ Method,”
International Journal of Heat and Mass Transfer, Vol. 44, pp. 3321-3333

Ding, H., Shu, C., and Yeo, K. S., 2005, “Simulation of Natural Convection in Eccentric Annuli
Between a Square Outer Cylinder and a Circular Inner Cylinder using Local MQ-DQ Method,”
Numerical Heat Transfer, Part A, Vol. 41, pp 291-313

Bishop, E. H., Carley, C. T. and Powe, R. E., 1968, “Natural Convective Oscillatory Flow in
Cylindrical Annuli,” International Journal of Heat and Mass Transfer, Vol. 11 pp. 1741-1752

Warrington, R. O., and Powe, R. E., 1984, “The Transfer of Heat by Natural Convection
Between Bodies and Their Enclosures,” International Journal of Heat and Mass Transfer, Vol.
28, No. 2, pp. 319-330

Lieberman, J. and Gebhart B., 1969, “Interactions in Natural Convection from and Array of
Heated Elements, Experimental,” International Journal of Heat and Mass Transfer, Vol. 12, pp.
1385-1396

Tokura, I., Saito, H., Kishinami, K. and Muramoto, K., 1983, “An experimental Study of Free
Convection Heat Transfer From a Horizontal Cylinder in a Vertical Array Set in Free Space
Between Parallel Walls,” ASME Journal of Heat Transfer, Vol. 105, pp. 102-107

Prusa, J. and Yao, L. S., 1983, “Natural Convection Heat Transfer Between Eccentric Horizontal

13
Cylinders,” ASME Journal of Heat Transfer, Vol. 105, pp. 108-116

Vasseur, P., Rodillard, L. and Chandra Shekar, B., 1983, “Natural Convection Heat Transfer of
Water Within a Horizontal Cylindrical Annulus with Density Inversion Effects,” ASME Journal
of Heat Transfer, Vol. 105, pp. 117-123

Moukalled, F. and Acharya, S., 1996, “Natural Convection in the Annulus Between Concentric
Horizontal Circular and Square Cylinders,” Journal of Thermophysics and Heat Transfer, Vol.
10, No. 3, July-Sept., pp. 524-531

Ghaddar, N. K., 1992, “Natural Convection Heat Transfer Between a Uniformly Heated
Cylindrical Element and its Rectangular Enclosure,” International Journal of heat and Mass
Transfer, Vol. 35, No. 10, pp. 2327-2334

Ekundayo, C. O., Probert, S. D. and Newborough, M., 1998, “Heat Transfer from a Horizontal
Cylinder in a Rectangular Enclosure,” Applied Energy, Vol. 61, pp. 57-78

Liu, Y., Phan-Thien, N. and Kemp, R., 1996, “Coupled Conduction-Convection Problem for a
Cylinder in an Enclosure,” Computational Mechanics, Vol. 18, pp. 429-443

Sasaguchi, K., Kuwabara, K., Kusano, K. and Kitagawa, H., 1998, “Transient Cooling of Water
Around a Cylinder in a Rectangular Caviti- A Numerical Analysis of the Effect of the Position of
the Cylinder,” International journal of Heat and Mass transfer, Vol. 41, pp. 3149-3156

Gebhart, B., Jaluria, Y., Mahajan, R. L. and Sammakia, B., 1988, Buoyancy-Induced Flows and
Transport, Hemisphere Publishing Corporation, New York, NY

Bahney, R. H. and Lotz, T. L., 1996, “Spent Nuclear Fuel Effective Thermal Conductivity
Report,” Prepared for the U.S. DOE, Yucca Mountain Site Characterization Project Office by
TRW Environmental Safety Systems, Inc. July 11, D.I.: BBA000000-01717-5705-00010 REV
00

Yang, K. T., and Lloyd, J. R., 1985, “Natural Convection-Radiation Interaction in Enclosures”
Natural Convection; Fundamentals and Applications, Edited by Kakac, S., Aung, W., and
Viskanta, R., Hemisphere Pubishing Corporation, pp. 381-410.

Lovett, P. M., 1991, An Experiment to Simulate the Heat Transfer Properties of a Dry,
Horizontal Spent Nuclear Fuel Assembly, submitted to the Department of Nuclear Engineering in
partial fulfillment of the requirements for the degree of Master of Science in Nuclear
Engineering at the Massachusetts Institute of Technology.

14
134.06 mm

Nitrogen Backfill

Zircaloy Channel
Tw = 298.15 K

UO2 pellet, Q [W]


134.06 mm

Zircaloy Cladding

Backfill gap, N2

Figure 1 Domain and grid for the 2D model

0.020

0.010

0.000

-0.010
φ /φ D

GRID A GRID B GRID C GRID D


-0.020

-0.030 % ∆Tc,max
% S,max
-0.040
% NuL

-0.050
10000 15000 20000 25000 30000 35000 40000
#Elements

Figure 2 Grid convergence of the 2D model

15
610 mm Zircaloy channel
walls, Tw=298.15 K
y Nitrogen Backfill
(inside)
x
z Uranium dioxide rod, Q[W]
134.06 mm

Stainless Steel end


plates, Tw = 298.15 K

Zircaloy cladding wall


134.06
Figure 3 Domain and grid for the 3D model

307
306 Elements

305 313280
304 163520

303 103320
T [K]

302
301
300
299
298
297
-0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.08
x [m]
Figure 4 Temperature profiles along horizontal line on center axial plane at 0.02 m above center

16
T [K] Simple 2D S [cm/s] Simple 2D

T [K] Full 2D S [cm/s] Full 2D

T [K] 3D S [cm/s] 3D

Figure 5 Temperature (T) and velocity magnitude (S) contours for Q = 5.2 W; comparison of the 2D models
with the 3D model.

17
T [K] S [cm/s]
z = -0.3 z = -0.3

z=0 z=0

z = 0.25 z = 0.25

x=0 x=0

V [cm/s] W [cm/s]
z = -0.3 z = -0.3

z=0 z=0

z = 0.25 z = 0.25

x=0 x=0

Figure 6 Temperature (T), velocity magnitude (S), y-velocity (V) and z-velocity (W) contours for Q = 5.2 W

30
2D Simple
25 2D Full
3D
20 2D Simple (Radiation)
T c-T w [K]

3D (Radiation)

15

10 Figure 7 Cladding to wall


temperature difference for
2D and 3D models with and
5 without radiation heat
transfer
0
0 1 2 3 4 5 6
Q [W]

18
14

12

10
Smax [cm/s] 8

6 2D Simple
2D Full
4
3D
2D Simple (Radiation)
2
3D (Radiation)
0
0 1 2 3 4 5 6
Q [W]

Figure 8 Maximum velocity magnitudes for 2D and 3D models with and without radiation heat transfer

1.2E+04
2D Simple
2D Full
1.0E+04
3D
2D Simple (Radiation)
8.0E+03
3D (Radiation)
GrD

6.0E+03

4.0E+03

2.0E+03

0.0E+00
0 1 2 3 4 5 6
Q [W]

Figure 9 Grashof number with respect to heat load

19
100
Reference Length, L = R0-Ri Ghaddar [1991]
90 Where R0 and Ri are radii of equivalent L/Ri=26.6
spheres with same volumes as the enclosure
80 and cylinder respectively
2D L/Ri=3.82
70
Simple
60 2D Full
2D
NuL

50

40 3D L/Ri=3.82
Warrington and
30 Powe [1984]
L/Ri=1.47
20
L/Ri=1.14
L/Ri=0.73
10
L/Ri=0.66
0
0.0E+00 1.0E+06 2.0E+06 3.0E+06 4.0E+06 5.0E+06 6.0E+06 7.0E+06
RaL
Figure 10 Comparison of results to previous studies

MATERIAL PROPERTIES

Nitrogen
ρ = 3.501 - 0.01362*T + 2.492E-5*T2 – 2.156E-8*T3 + 7.101E-12*T4 [kg/m3];
Data from Bahney and Lotz [1996]
Cp(T) [j/kgK]; default polynomial in Fluent 6.2 Database
k(T) [W/mK]; default polynomial in Fluent 6.2 Database
µ = 1.663E-5 [kg/ms]

Uranium Dioxide (Bahney and Lotz, 1996)


ρ = 10970 [kg/m3]
Cp = 247 [j/kgK]
k = 8.36 [W/mK]
ε = 0.79

Zircaloy
ρ = 6500 [kg/m3]
Cp = 350 [j/kgK]
k = 12.58 [W/mK]
ε = 0.806

Stainless Steel (Bahney and Lotz, 1996)


ρ = 8000 [kg/m3]
Cp = 481.68 [j/kgK]
k = 13.3 [W/mK]
ε = 0.6

20

Вам также может понравиться