Вы находитесь на странице: 1из 105

Evaluation of sub-channel ow mixing coecient for

typical PWR fuel bundles having spacers using CFD


analysis
by
Vinay Karanam
Enrolment No. - ENGG01200801059
Bhabha Atomic Research Centre, Mumbai
A dissertation submitted
to the Board of Studies in Engineering Sciences
In partial fulllment of requirements
For the Degree of
Master of Technology
HOMI BHABHA NATIONAL INSTITUTE
JULY 2011
HOMI BHABHA NATIONAL INSTITUTE
Recommendations of the Thesis Examining Committee
As members of the Thesis examining Committee, we recommend that the dis-
sertation prepared by Vinay Karanam entitled Evaluation of sub-channel ow
mixing coecient for typical PWR fuel bundles having spacers using CFD analy-
sis be accepted as fullling the dissertation requirement for the Degree of Master
of Science(Engineering).
Name Signature
Member-1 Mr. P.V.Durgarasad
Member-2 Mr. P.Majumdar
Member-3 Mr. M.Madhusoodanan
Technical Advisor Mr. Joe Mohan
Co-guide Dr. J.S.Jayakumar
Guide/Convener Dr. N.K.Maheshwari
Chairman Dr. G.R.Reddy
Final approval and acceptance of this dissertation in contingent upon the
candidates submission of the nal copies of the dissertation to HBNI.
Date:
Place:
i
DECLARATION
I, hereby declare that the investigation presented in the thesis has been carried
out by me. The work is original and has not been submitted earlier as a whole
or in part for a degree / diploma at this or any other Institution / University.
Vinay Karanam
Reactor Projects Division
Bhabha Atomic Research Centre, Mumbai
July, 2011
ii
To my Parents ...
iii
ACKNOWLEDGMENT
I express my sincere gratitude to my project guides Dr. J.S.Jayakumar and Dr.
N.K.Maheshwari for their valuable guidance, timely suggestions and constant
encouragement. His interest and condence in me has helped immensely for the
successful completion of this work. I am also thankful to Sh.R.Dinesh Babu,
my section head and Sh. Joe Mohan, my technology advisor for their valuable
time to time suggestion and their involvement in the project. I am thankful to
Inder Kumar, Chirag Patel, Lt.Cdr. Prakash, Ms. Sugandhi, B.K.Tripathy for
their suggestions, generous help and for maintaining friendly environment around
me.
Special thanks to Sh.K.N.Vyas for giving me this opportunity to work on
this project. I would like to special thanks to Sh.R.R.S.Yadav for his help,
guidance and suggestions for the development of this project .
I would also like to thank Gowri, Sridhar and Lovely for making my stay
at B.A.R.C. Training School Hostel, memorable and enjoyable and making me
feel the importance of good friends. Special thanks to my B.Tech teacher Prof.
D.Varadaraju whose constant support and encouragement led me to have this
great opportunity of joining B.A.R.C., Mumbai.
Finally, I would like to thank my parents, my brother, in-laws for their con-
stant support and encouragement. A very special thanks to my caring wife Kriti
for being with me unconditionally.
Vinay Karanam
Abstract
The investigation of thermal-hydraulic characteristics of rod bundles is impor-
tant for safe operation and optimal design of nuclear reactor power plants. For
better utilization of fuel, the thermal margin on Critical Heat Flux (CHF) and
fuel centerline temperature must be judiciously determined. CHF is of interest
from consideration of maximum permissible plant power level. Uncertainties in
the CHF evaluation will lead to higher factor of safety resulting in de-rating of
plant power. Core thermal hydraulics is generally carried out using sub-channel
analysis codes such as COBRA. This sub-channel analysis code requires values of
turbulent mixing factor as input. For correct predictions of fuel performance,
accurate value of turbulent mixing factor is essential. The turbulent mixing
factor is highly dependent on geometry and Reynolds number. To estimate
the value of , experiment must be carried out for each of the geometries under
study. This in general is very laborious, time consuming and expensive. In this
paper, a methodology to estimate the value of using CFD modeling is pre-
sented. CFD can be used to simulate typical experimental condition and help
in estimation of these coecients without necessity of having elaborate experi-
mental study. This method has the advantage that complex spacer geometries,
which are not easily amenable to experimental determination, can also be mod-
eled. The paper describes the native method of estimating the turbulent mixing
coecient from the CFD results. The methodology of estimation of using CFD
data is rst benchmarked against experimental results. For this, two cases have
been analyzed, (i) isothermal ow through triangular lattice of bare fuel pins,
and (ii) non-uniform heating in a square lattice pins with egg crate spacer grids.
The computational results are found to be matching well with the experimental
results within margin of experimental uncertainty. After extensive validation of
the methodology, it has been applied to the actual fuel assembly under study.
By judicious engineering approximation, one can single out a representative for
further analysis i.e. input into the sub-channel analysis code, such as COBRA.
The paper also describes results of the subchannel analysis.
vi
Contents
Certicate ii
Declaration iii
Acknowledgements iv
Abstract v
List of Figures xi
List of Tables xvi
Nomenclature xix
1 INTRODUCTION AND OBJECTIVE 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Thesis Organisation . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2 LITERATURE SURVEY 4
2.1 Prior Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Experimental Work . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.3 Computational Work . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.4 Comments on the work done so far . . . . . . . . . . . . . . . . . 15
vii
2.5 Closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3 COBRA CODE and GOVERNING EQUATIONS OF CFD 17
3.1 Introduction to COBRA . . . . . . . . . . . . . . . . . . . . . . . 17
3.1.1 Basic Equations . . . . . . . . . . . . . . . . . . . . . . . . 18
3.1.2 Single-Phase Turbulent Mixing . . . . . . . . . . . . . . . 19
3.2 Introduction to CFD and turbulence . . . . . . . . . . . . . . . . 20
3.2.1 Reynolds Average Navier Stokes Simulation (RANS) . . . 20
3.2.2 Eddy-Viscosity Models (EVM) . . . . . . . . . . . . . . . . 21
3.2.3 Reynolds-Stress Models (RSM) . . . . . . . . . . . . . . . 22
3.3 Closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4 ANALYSIS AND VALIDATION WITH THE EXPERIMENTS 23
4.1 Introduction to Analysis . . . . . . . . . . . . . . . . . . . . . . . 23
4.2 Description of Methodology . . . . . . . . . . . . . . . . . . . . . 23
4.2.1 Grid Preparation . . . . . . . . . . . . . . . . . . . . . . . 23
4.2.2 Choice of Turbulence Model . . . . . . . . . . . . . . . . . 24
4.2.3 Plotting vs. Axial length . . . . . . . . . . . . . . . . . . 25
4.3 Case (i): Triangular Lattice - Isothermal Case, Roidt et. al.[1] . . 27
4.3.1 Description of Experimental set-up . . . . . . . . . . . . . 27
4.3.2 Description of Model and CFD input parameters . . . . . 29
4.3.3 Results and comparison with the experiment . . . . . . . . 30
4.3.4 Grid sensitivity of the model . . . . . . . . . . . . . . . . . 34
4.3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.4 Case (ii): Square Lattice with egg crate spacer grids (four Nos.)
with Non-uniform Heating . . . . . . . . . . . . . . . . . . . . . . 36
4.4.1 Description of Experimental set-up . . . . . . . . . . . . . 36
4.4.2 Description of Model and CFD input parameters . . . . . 39
viii
4.4.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . 40
4.4.4 Grid sensitivity of the model . . . . . . . . . . . . . . . . . 44
4.4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.5 Representative Models . . . . . . . . . . . . . . . . . . . . . . . . 49
4.5.1 Square Lattice - Isothermal Case . . . . . . . . . . . . . . 49
4.5.2 Square Lattice - Uniform heating Case . . . . . . . . . . . 51
4.5.3 Square Lattice - Non-uniform heating Case . . . . . . . . . 51
5 PREDICTION OF TURBULENCE MIXING BEHAVIOUR US-
ING CFD FOR SPACER LOCATION IN A TYPICAL PWR 55
5.1 Description of problem under study . . . . . . . . . . . . . . . . . 55
5.2 Description of CFD models . . . . . . . . . . . . . . . . . . . . . 56
5.2.1 Bare model . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.2.2 With one spacer . . . . . . . . . . . . . . . . . . . . . . . . 57
5.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . 57
5.4 Grid sensitivity of the model . . . . . . . . . . . . . . . . . . . . . 63
5.4.1 Model Type-1 . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.4.2 Model Type-2 . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.6 Closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
6 SUBCHANNEL ANALYSIS OF A TYPICAL PWR USING THE
SUBCHANNEL CODE COBRA-IV 69
6.1 Subchannel Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.1.1 Description of input le . . . . . . . . . . . . . . . . . . . 69
6.1.2 Results and Discussion . . . . . . . . . . . . . . . . . . . . 70
6.1.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.2 CFD Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
ix
6.2.1 Description of CFD model . . . . . . . . . . . . . . . . . . 72
6.2.2 Results and Comparison with the COBRA-IV results . . . 73
6.2.3 Grid sensitivity of the model . . . . . . . . . . . . . . . . . 74
6.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.4 Closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
7 CONCLUSIONS AND SCOPE FOR THE FUTURE WORK 78
7.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
7.2 Scope of the Future Work . . . . . . . . . . . . . . . . . . . . . . 79
References 80
x
List of Figures
2.1 Summary of the Experimental Conditions of the Referenced Pre-
vious Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Various experimental results for mixing coecient . . . . . . . . 9
4.1 Subchannel connection . . . . . . . . . . . . . . . . . . . . . . . . 26
4.2 Cell conguration in the plane extracted. . . . . . . . . . . . . . . 26
4.3 A Typical Plot showing vs. Axial length of the plane . . . . . . 26
4.4 Cross section of the test section of the experimental set-up . . . . 28
4.5 Schematic Layout of the Duct in the experimental set-up. . . . . . 29
4.6 Cross sectional view of the model case(i) . . . . . . . . . . . . . . 29
4.7 Isometric view of the model case(i) . . . . . . . . . . . . . . . . . 30
4.8 Ethane concentration in the downstream position [1] . . . . . . . 32
4.9 Ethane concentration in the upstream position [1] . . . . . . . . . 32
4.10 The Concentration gradient across the clearance gap [1] . . . . . . 33
4.11 Normal Velocity through the axial length across the clearance gap 33
4.12 Plot of Vs. Axial Length for case (i) . . . . . . . . . . . . . . . 34
4.13 Cross section of the very coarse mesh (Mesh 1) conguration of
case(i) model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.14 Cross section of the coarse mesh (Mesh 2) conguration of case(i)
model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.15 Cross section of the ne mesh (Mesh 3)conguration of case(i) model 36
xi
4.16 Cross section of the very ne mesh (Mesh 4) conguration of case(i)
model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.17 A cross-section of the test geometry . . . . . . . . . . . . . . . . . 38
4.18 Schematic sketch of the experimental test set-up . . . . . . . . . . 38
4.19 Cross sectional view of the model case(ii) . . . . . . . . . . . . . . 39
4.20 A isometric view of the model case(ii) . . . . . . . . . . . . . . . . 40
4.21 Plot of vs. error s . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.22 Plot of vs. rows of the intersecting plane for a bare model . . . 42
4.23 Plot of vs. rows of the intersecting plane for the model with
spacers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.24 Plot of vs. row number of the intersecting plane showing details
near rst and second spacer . . . . . . . . . . . . . . . . . . . . . 43
4.25 Plot of vs. row number of the intersecting plane showing details
near rst spacer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.26 Cross section of the very coarse mesh (Mesh 1) conguration of
case(ii) bare model . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.27 Cross section of the coarse mesh (Mesh 2) conguration of case(ii)
bare model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.28 Cross section of the ne mesh (Mesh 3)conguration of case(ii)
bare model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.29 Cross section of the very ne mesh (Mesh 4) conguration of
case(ii) bare model . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.30 Cross section of the very coarse mesh (Mesh 1) conguration of
case(ii) model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.31 Cross section of the coarse mesh (Mesh 2) conguration of case(ii)
model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.32 Cross section of the ne mesh (Mesh 3)conguration of case(ii) model 48
xii
4.33 Cross section of the very ne mesh (Mesh 4) conguration of
case(ii) model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.34 Plot of vs. Axial Length of the intersecting plane with one spacer
grid - isothermal case . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.35 Plot of vs. Axial Length of the intersecting plane with one spacer
grid -uniform heating . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.36 Plot of vs. Axial Length of the intersecting plane with one spacer
grid - non-uniform heating . . . . . . . . . . . . . . . . . . . . . . 53
4.37 Comparison of values of isothermal case, uniform heating, non-
uniform heating and bare model . . . . . . . . . . . . . . . . . . . 53
5.1 Cross section of the channel with the spacer grid . . . . . . . . . . 56
5.2 Bare model of the problem under study . . . . . . . . . . . . . . . 56
5.3 Cross sectional view of the model type-1 . . . . . . . . . . . . . . 57
5.4 Cross sectional view of the model type-2 . . . . . . . . . . . . . . 58
5.5 Isometric view of the model type-1 . . . . . . . . . . . . . . . . . 58
5.6 Isometric view of the model type-2 . . . . . . . . . . . . . . . . . 59
5.7 The plot againt row numbers along axial direction for model
type-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.8 The plot againt row numbers along axial direction for model
type-2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.9 Vortex at the entry of spacer wall . . . . . . . . . . . . . . . . . . 61
5.10 Vortex at the exit of spacer wall . . . . . . . . . . . . . . . . . . . 61
5.11 Helical movement of the coolant . . . . . . . . . . . . . . . . . . . 62
5.12 Vortex downstream the spacer . . . . . . . . . . . . . . . . . . . . 62
5.13 The ow changing direction downstream the spacer . . . . . . . . 63
xiii
5.14 Cross section of the very coarse mesh (Mesh 1) conguration of
model type-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.15 Cross section of the coarse mesh (Mesh 2) conguration of model
type-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.16 Cross section of the ne mesh (Mesh 3)conguration of model type-1 65
5.17 Cross section of the very ne mesh (Mesh 4) conguration of model
type-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.18 Cross section of the very coarse mesh (Mesh 1) conguration of
model type-2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.19 Cross section of the coarse mesh (Mesh 2) conguration of model
type-2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.20 Cross section of the ne mesh (Mesh 3)conguration of model type-2 67
5.21 Cross section of the very ne mesh (Mesh 4) conguration of model
type-2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.1 A cross sectional view of the bare model of 1/12th symmetric chan-
nel for analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.2 A cross sectional view of the model with one spacer of 1/12th
symmetric channel for analysis . . . . . . . . . . . . . . . . . . . . 73
6.3 A isometric view of the model with one spacer of 1/12th symmetric
channel for analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.4 Cross section of the very coarse mesh (Mesh 1) conguration of
1/12th symmetric model with one spacer for analysis . . . . . . . 76
6.5 Cross section of the coarse mesh (Mesh 2) conguration of 1/12th
symmetric model with one spacer for analysis . . . . . . . . . . . 76
6.6 Cross section of the ne mesh (Mesh 3) conguration of 1/12th
symmetric model with one spacer for analysis . . . . . . . . . . . 77
xiv
6.7 Cross section of the very ne mesh (Mesh 4) conguration of 1/12th
symmetric model with one spacer for analysis . . . . . . . . . . . 77
xv
List of Tables
2.1 List of a few values reported in literature . . . . . . . . . . . . . 6
4.1 Results of various grid conguration to assesss the optimize the
mesh size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2 Results of various Turbulence model to obtain . . . . . . . . . . 25
4.3 The coecients of RSM turbulence model . . . . . . . . . . . . . 30
4.4 Input Parameters for the case(i)[1] . . . . . . . . . . . . . . . . . 31
4.5 Results of obtained from the experimental set-up [1] . . . . . . 31
4.6 vs. various grid conguration of case(i) . . . . . . . . . . . . . . 35
4.7 Comparison between the result of obtained by CFD and experi-
mental work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.8 Input Parameters for the case(ii) [2] . . . . . . . . . . . . . . . . . 40
4.9 vs. various grid conguration of case(ii) bare model . . . . . . . 44
4.10 vs. various grid conguration of case(ii) . . . . . . . . . . . . . 46
4.11 Input Parameters of the CFD for the isothermal case of represen-
tative model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.12 vs. various grid conguration of representative model (isothermal) 50
4.13 Input Parameters of the CFD for the uniform heating case of rep-
resentative model . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.14 Input Parameters of the CFD for the non-uniform heating case of
representative model . . . . . . . . . . . . . . . . . . . . . . . . . 52
xvi
5.1 vs. various grid conguration of model type-1 . . . . . . . . . . 63
5.2 vs. various grid conguration of model type-2 . . . . . . . . . . 66
6.1 Average temperature at the exit and CHF margin of 1/12th sym-
metric model with ten spacer grids by COBRA analysis with value
of beta given in the literature . . . . . . . . . . . . . . . . . . . . 70
6.2 Average temperature at the exit and CHF margin of 1/12th sym-
metric model with ten spacer grids by COBRA analysis with beta
obtained from CFD . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.3 Average temperature at the exit and CHF margin of 1/12th sym-
metric bare model by COBRA analysis with beta obtained from
CFD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.4 Comparison table showing average temperature at the exit given
by CFD analysis and COBRA analysis . . . . . . . . . . . . . . . 75
6.5 Average temperature at the exit vs. various grid conguration of
1/12th symmetric model with one spacer for analysis . . . . . . . 75
xvii
xviii
Nomenclature
CFD Computational Fluid Dynamics
PWR Pressurized Water Reactor
RSM Reynolds Stress Model
SST Shear Stress Model
A

Average Area
(A
j
+A
j1
)
2
(m
2
)
[DC] Dierence Operator for subchannels
[DC]
T
Transpose of Dierence Operator
G Mass Velocity (Kg/sec m
2
)
m Flow rate (Kg/sec)
L Channel Length (m)
P
w
Wetted Perimeter (m)
u, v, w Velocity components in X, Y and Z directions respectively (m/s)
V
n
Velocity component in normal direction to the plane(m/s)
U, V, W Reynolds averaged velocity components in X, Y and Z directions respectively (m/s)
p Pressure (N/m
2
)
P Reynolds averaged pressure (N/m
2
)
k Turbulent kinetic energy (m
2
/s
2
)
P Production rate of turbulent kinetic energy (m
2
/s
3
)
Dissipation rate of turbulent kinetic energy (m
2
/s
3
)
Specic dissipation rate of turbulent kinetic energy (1/s)
v
2
Wall normal component of the Reynolds stress tensor (m
2
/s
2
)
f Scalar representing redistributive source term in the v
2
equation (1/s)
V, L, T Velocity, length and time scales of the uctuating quantities respectively
Density (kg/m
3
)
Molecular viscosity (kg/m s)

t
Turbulent or Eddy viscosity (kg/m s)
Turbulence Mixing Coecient (dimensionless)

+
Normalised dissipation rate of turbulent kinetic energy
_


u
4

+
t
Normalised kinematic eddy viscosity
_

_
Shear stress (N/m
2
)

w
Wall shear stress
__

_
U
y
_
y=0
_
A Surface area vector
F
f
Volume ux
F
df
Diusion ux of variable
xix
Chapter 1
INTRODUCTION AND
OBJECTIVE
1.1 Introduction
In most of nuclear reactors, fuel rod bundles are the essential elements that con-
sist of tightly packed rod arrays. These fuel rods are surrounded by the owing
coolant which plays a signicant role in heat removal. Therefore, the investigation
of thermal-hydraulic characteristics inside the rod bundles is important for opti-
mal design and safety operation of a nuclear reactor power plant. Spacer grids are
one of the main components used on the rod bundles in order to maintain appro-
priate rod-to-rod clearance, secure ow passage, and prevent the bundle damage
from ow-induced vibration(FIV), etc. Existence of the grids would signicantly
inuence the ow and temperature distributions within the rod bundles.
Numerous experiments in the past (Trupp and Azad (1975)[4]; Vonka. V
(1988)[7]; Wu and Trupp (1993) have clearly revealed that ow conditions inside
fuel bundles are very dierent from those typical of pipe ow. The near wall
turbulence anisotropy in noncircular ducts causes the formation of secondary
vortices inside the channel, which lead the coolant to actually spiral through the
bundle. Classical codes used in nuclear industries, such as sub-channel analysis
codes COBRA do not consider these phenomena and the value it takes as input
based on certain correlations such as Rogers and Rosenhart (1972), are not ap-
propriate for this purpose, being calibrated on specic geometries and operational
conditions. The application of an adequate and general numerical methodology is
necessary, which should provide an accurate value which would represent those
1
phenomena and help in sub-channel analysis for general fuel arrangements.
Computational uid dynamics (CFD) methodology then attracts more atten-
tion from the researcher to model these complicated phenomena, including the
swirling secondary ow, turbulent mixing and heat transfer enhancement down-
stream a spacer with appropriated vanes. Lee and Jang(1997)[12] performed
numerical simulations for a rod bundle using a nonlinear k model without ad-
justment and concluded that this approach strongly underestimated the strong az-
imuthally turbulence intensity. However, by adjusting model coecients adopted
in a quadratic k (Shih et al. (1993)[23]), Baglietto and Ninokata (2005)[15]
had shown its promising capability inadequate anisotropy modeling of the wall-
shear stress distribution and the velocity eld in tight lattice fuel bundles. Kazuo
Ikeda (2005)[16], Cheng and Tak (2006)[18], C.C. Liu and Y.M. Ferng (2009)[21],
Toth and Aszodi (2010)[17] have successfully used CFD to predict CHF, ow and
temperature distribution.
1.2 Objective
The thermal hydraulic characteristic of bare fuel rod ow can be dierent from
the spacer lled fuel rod ow. Some of the key phenomena which are dicult to
model but are signicantly important to assess the spacer lled fuel rod ow is
turbulence mixing behavior, secondary ow etc. Hence, the system code needs to
be benchmarked against experimental data of such phenomena before applying
them to the new generation reactors. The main objective of this thesis work is
to:
1. Evaluate the applicability of CFD code to predict the turbulence mixing
behavior of spacer locations by comparing the experimental data generated in a
few test facilities referred in the literature with the code predictions.
2. Applying this methodology on a typical PWR to estimate the turbulence
mixing coecient . The value thus obtained is input into COBRA code to
predict the temperature distribution in the channel. This temperature prole ob-
tained by COBRA code is veried by the CFD code predictions.
2
1.3 Thesis Organisation
This dissertation is divided in total 7 Chapters.
Chapter 2 summarizes the work documented in literature for applicability of
CFD code to simulate the turbulence mixing of ow in fuel rod bundles and other
experimental and analytical works in estimation of turbulence mixing coecients.
Chapter 3 gives brief description about COBRA code and discusses the turbu-
lence models treated in the thesis and the governing equations for the RANS.
Chapter 4 gives brief description about analysis and validation of methodology
with the experimental data available in the literature.
Chapter 5 details the prediction of turbulence mixing behavior using CFD for
a typical PWR with spacer grids.
Chapter 6 discusses the ndings in the work by incorporating results of CFD
code prediction in COBRA code.
Chapter 7 summarizes the salient concluding features of project work.
3
Chapter 2
LITERATURE SURVEY
This chapter is devoted to the summary of the literature study. The comments
on the work done so far and motivation for this project is also detailed.
2.1 Prior Work
The scientic community has been interested in estimating the turbulence mixing
coecient ever since commercial scale reactors have started to come at rapid pace.
However, these studies were more in experimental in nature and there are limited
studies on the applicability of CFD code to simulate turbulence mixing coecient.
A review of literature is mentioned below:
2.2 Experimental Work
Roidt et al. (1970)[1] studied the turbulence mixing phenomena on bare rods
in single phase regimes using a tracer gas. A tracer gas is injected into a single
subchannel of a large air ow model of a reactor rod bundle. The tracer uxes
are sampled at two positions downstream to obtain the axial variations in the
uxes through both the injection subchannel and those adjacent to it. Combining
this information with measured subchannel area changes (which allow estimating
the axial change in volumetric ow rates) yields measures of the turbulent eddy
diusion coecient and subchannel cross ows. The results agree with other
investigations. Turbulent exchange coecients calculated from the data vary
by 50 percent, variations which should be expected when correlating exchange
coecients from experiments with dierent internal distributions of transportable
4
scalar quantities.
Castellana (1973)[2] had estimated Single-phase sub-channel mixing data, ob-
tained from a 25 rod square array by measuring precisely sub-channel exit tem-
peratures over a range of test conditions. A least-squares type statistic operating
on exit enthalpy dierences was developed to ascertain, in conjunction with the
COBRA-II sub-channel computer analysis, an optimum value for the coecient of
Rowe and Angless single-phase mixing correlation of . When the same analyti-
cal procedures were applied to data taken under conditions of subcooled nucleate
boiling, an approximately linear correlation of with average exit quality was
found. The values obtained were for conditions of natural, or unaided mixing,
and for design purposes should be considered as a dependable lower limit. They
concluded that commercial rod-spacer designs intended to increase mixing will,
of course, show substantially higher values of .
Seale W. J. (1977)[3] had estimated mixing by turbulent diusion and sec-
ondary ow between the parallel subchannels of ducts simulating smooth, bare
rod bundles has been investigated both experimentally and analytically. The
experimental work was performed on a long wind tunnel which was designed to
achieve an axially unchanging, fully developed temperature prole at the tunnel
exit. This is believed to be the rst substantial investigation of inter-subchannel
mixing using this technique. Detailed measurements were made over a range of
Reynolds numbers and in three congurations simulating pitch to diameter ratios
of 1.833, 1.375 and 1.1. The results of the experimental work conrm the major
ndings of previous investigations, in particular that the inter subchannel mixing
rates are considerably higher than predicted by simple diusion theory, and are
relatively insensitive to variations in the gap width between the rods. Eective
diusivities through the gap appear to be strongly anisotropic and there is no
evidence of secondary ows. The analytical work has been performed by solv-
ing the basic dierential equations of the turbulent ow and heat ux using the
k turbulence model and the Gosman numerical integration procedure.
This is believed to be the rst application of these procedures to the prediction
of inter-subchannel mixing. A detailed comparison between the computer predic-
tions and the results of the experimental work revealed that secondary ows were
not an important determinant in the mixing rate and did not allow agreement to
be obtained with the measured axial velocity and temperature contours. Predic-
5
tions assuming that the eective diusivity was isotropic, produced gap Stanton
numbers an order less than those measured. To obtain gap stanton numbers of
the correct order it was found necessary to use anisotropic eective diusivities in
which the anisotropy distribution rose sharply near the walls to very high values
(approx. 50).
Trupp and Azad (1974)[4] had carried out a wind tunnel study of fully de-
veloped uniform-density turbulent ow through triangular array rod bundles is
described. Measurements were made for three tube spacings (P/D = 1.20, 1.35
and 1.50) over a Reynolds number range of 12000-84000. The data include fric-
tion factors, local wall shear stresses, and the distributions of mean axial velocity,
Reynolds stresses and eddy diusivities. The secondary ow pattern is inferred
from the available evidence.
The list of a few values reported in literature is shown in the table 2.2.
Experimenter Channel Fluid S/d D
h
Experimental
Type (mm) technique
Rowe T-T water 0.036 5.105 exit enthalpy 0.063 Re
0.1
and Angle and code simulation
Castellana S-S water 0.334 13.56 exit enthalpy 0.021 Re
0.1
and code simulation
Seale S-S air 0.1 27.1 pitot temperature 0.029 Re
0.1
0.375 57.3 probe for 0.016 Re
0.1
0.833 125 temperature and 0.009 Re
0.1
velocity distribution
Rogers and S-S - - - - 0.004 (D
h
/S) Re
0.1
Rosehart
Rogers and T-T water 0.4 25.4 Tracer and 0.005 (D
h
/S)
Tahir Pressure balance (S/d)
0.106
Re
0.1
Roidt T-T air 0.256 31.5 Tracer and
calculated diversion
cross plains
Table 2.1: List of a few values reported in literature
The experimental conditions of various researches in the literature is summa-
rized in the Fig 2.1.
6
Figure 2.1: Summary of the Experimental Conditions of the Referenced Previous
Studies
Krauss and Meyer (1997)[5] investigated the turbulent air ow in a central
channel of heated 37-rod bundles with triangular array at two dierent pitch-to-
diameter ratios (P/D=1.12 and P/D=1.06). Measurements were performed with
a hot-wire probe with x-wires and an additional temperature wire. Time mean ve-
locities, time mean uid temperatures, wall shear stresses and wall temperatures,
7
turbulent quantities such as the turbulent kinetic energy, all Reynolds stresses
and all turbulent heat uxes were measured at two dierent pitch-to-diameter
ratios in a central channel of the bundle. It is shown that with decreasing gap
width the turbulence eld in rod bundles deviates signicantly from that in a
circular tube. Also, data on the power spectral density functions of the velocity
and temperature uctuations are presented. These data show the existence of
large-scale periodic uctuations of velocity and temperature in the gap region of
two adjacent rods. These uctuations are responsible for the high inter subchan-
nel heat and momentum exchange. Spectral measurements with two hot wire
probes imply a distinct similarity of motion of vortices in adjacent subchannels
of the bundle.
Creer et.al (1978) [6] performed experiment on and found local velocity and
turbulence intensity with a laser Doppler anemometer near ow blockages in an
unheated 7x7 rod bundle. Sleeve blockages were positioned on the center nine
rods to create area reductions of 70 and 90 percent in the center four subchannels
of the bundle. Experimental results indicated that extensive ow disturbances
existed downstream from the blockage clusters and mowed that only minor dis-
turbance can be expected upstream from the blockages. Recirculation zones for
both 70 and 90 percent blockages were detected downstream from the block-
age clusters and persisted for approximately three to ve subchannel hydraulic
diameters, depending on the degree of the blockage. The experimental velocity
results obtained with blockage clusters located midway between grid spacers were
successfully predicted using the COBRA subchannel computer program.
Vonka (1987)[7] practically observed secondary ow vortices in innite rod
bundles which have been predicted by a number of theoretical analyses has been
experimentally veried and noted that the magnitude of the secondary velocities
appeared to be less than the accuracy of the experimental techniques used. Only
indications of the maximum velocity magnitude have been available but no re-
port on successful direct measurement is known to the author. At ECN, laser
Doppler velocimetry is successfully used for measurement of secondary ow vor-
tices in two regular subchannels of a triangularly arranged bare rod bundle with
pitch-to-diameter ratio P/D = 1.3 under the Reynolds number conditions 60000
and 175000. One single secondary vortex, having the average tangential velocity
8
slightly less than 0.1 percent of the mean bulk velocity, is resolved per minimum
symmetry sector of the bundle geometry. Ensemble averages are made to obtain
quantitative description of the vortex and to form a data base for comparison
with calculations.
Elvis E. Dominguez-Ontiveros and Yassin A. Hassan (2009)[8] studied and de-
veloped an experimental data base with high spatial and temporal resolution of
uid ow velocity inside a 55 rod bundles with spacer grids. The full-eld detailed
data base is intended to validate CFD codes at various temporal-spatial scales.
Measurements are carried out using dynamic particle image velocimetry (DPIV)
technique inside an optically transparent rod bundle utilizing the matching index
of refraction (MIR) approach. This work presents full eld velocity vectors and
turbulence statistics for a rod bundle under single phase ow conditions.
Various experimental results for mixing coecient is shown in the gure 2.2.
Figure 2.2: Various experimental results for mixing coecient
2.3 Computational Work
Ramm et. al.(1973) [9] had done a theoretical analysis to study molecular and
turbulent transport phenomena between subchannels of innite bare rod arrays
at laminar, transition and turbulent ow conditions. For this investigation, the
theoretical approach of Ramm and Johannsen for predicting turbulent momentum
9
and heat transfer in rod bundles has been extended to evaluate three-dimensional
temperature elds. Results are presented enabling the prediction of the onset
and growth of laminarization in typical subchannels of square and triangular rod
arrays. These results are further applied to interpret the characteristic eects of
variations in Reynolds number, Prandtl number or geometric spacing on integral
exchange parameters as the thermal mixing ow rate and mixing length scale.
These results are of particular signicance relative to the explanation of recent
data from tracer-type mixing experiments and also exhibit the importance of
secondary ow eects on turbulent intersubchannel energy transport, in view of
these ndings, the physical relevance of current correlations derived from integral-
type experiments to numerically predict exchange coecients for use in lumped
parameter subchannel analysis codes is discussed.
Rapley and Gosman (1986)[10] found results from the application of a nite-
volume calculation method to fully-developed axial turbulent ow in various
smooth rod bundle arrangements. Simplied algebraic versions of the Reynolds
stress transport equations are used in the calculation of the full three dimensional
velocity eld, without any special adjustments for each geometry. The predic-
tions obtained for dierent rod spacings compare favourably with experiment and
reveal the signicant role of the cross-plane turbulence-driven secondary ow in
shaping the mean ow and turbulence distributions. The success of the results
obtained establish the eectiveness of the method and encourage further appli-
cations and development.
A.K. Mohanty and K.M. Sahoo (1987)[11] studied turbulent ow through rod
bundle sub-channels and analysed using a two-dimensional eddy diusivity and a
higher order K-l model. The eddy diusivity is a generalization of the Reichardts
model for a concentric annulus. The K-I model is used to examine the eects of
anisotropic turbulent diusion. Fully developed ow and thermal conditions are
assumed with the rod surface subjected to uniform heat ux. Both 30o and 45o
sub-channels, appropriate respectively for triangular and square arrangements of
rods, are investigated. The pitch to diameter ratio (PDR) was varied from 1.1 to
1.5. The algebraic eddy diusivity is found to reproduce literature information,
including those ascribable to anisotropy or secondary ow eects. Thermal cal-
culations have been carried out using Pr = 0.7 and both constant and variable
values of turbulent Prandtl number. Calibrating calculations are performed for
10
the concentric annular geometry.
Lee and Jang (1996)[12] studied the fully developed turbulent ows in closely
spaced bare rod arrays have been studied numerically by using a nonlinear k
model. This model permits inequality of the Reynolds normal stresses, a neces-
sary condition for calculating turbulence-driven secondary ows in non-circular
ducts. Results of the calculation are compared with experimental data. The
present computations yield reasonable agreement with experiments. However,
the modeling of the transport eect by large scale cross gap eddy motion is re-
quired.
Sadatomi et. al. (1995)[13] had presented a simple method for predicting the
single-phase turbulent mixing rate between adjacent subchannels in nuclear fuel
bundles. In this method, the mixing rate is computed as the sum of the two com-
ponents of turbulent diusion and convective transfer. Of these, the turbulent
diusion component is calculated using a newly dened subchannel geometry fac-
tor F* and the mean turbulent diusivity for each subchannel which is computed
from Elders equation. The convective transfer component is evaluated from a
mixing Stanton number correlation obtained empirically in this study. In order to
conrm the validity of the proposed method, experimental data on turbulent mix-
ing rate were obtained using a tracer technique under adiabatic conditions with
three test channels, each consisting of two subchannels. The range of Reynolds
number covered was 5000-66 000. From comparisons of the predicted turbulent
mixing rates with the experimental data of other investigators as well as the au-
thors, it has been conrmed that the proposed method can predict the data in a
range of gap clearance to rod diameter ratio of 0.02-0.4 within about 25 percent
for square array bundles and about 35 percent for triangular array bundles.
Hae-Yong Jeong et. al. (2006)[14] examined existing experimental data re-
lated to the turbulent mixing factor in rod arrays and a new denition of the
turbulent mixing factor is introduced to take into account the turbulent mixing
of uids with various Prandtl numbers. The new denition of the mixing factor is
based on the eddy diusivity of energy. With this denition of the mixing factor,
it was found that the geometrical parameter,
i
j /Dh, correlates the turbulent
mixing data better than S/d, which has been used frequently in existing correla-
tions. Based on the experimental data for a highly turbulent condition in square
rod arrays, a correlation describing turbulent mixing dependent on the parameter
11

i
j/Dh has been developed. The correlation is insensitive to the Re number and
it takes into account the eect of the turbulent Prandtl number. The proposed
correlation predicts a reasonable mixing even at a lower S/d ratio.
E. Baglietto and H. Ninokata (2004)[15] studied performances of various tur-
bulence models and evaluated calculation of detailed coolant velocity distribution
in a tight lattice fuel bundle. The individual models are briey outlined and com-
pared with respect to the prediction of wall shear stress and velocity eld, for a
fully developed ow inside a triangular lattice bundle. Comparisons clearly show
the importance of proper modeling of the turbulence-driven secondary ows in
subchannels. A quadratic k model, which showed promising capability in this
respect, is adjusted in its coecients, and the adjusted model is applied to fully
developed ow in an innite triangular array, with various Reynolds numbers.
The results show that the inclusion of adequate anisotropy modeling enables to
accurately reproduce the wall shear stress distribution and velocity eld in tight
lattice fuel bundles.
Kazuo Ikeda et. al. (2005)[16] made a numerical study using computational
uid dynamics (CFD) was carried out to estimate pressure loss in strap and
mixing vane structures. Moreover, a CFD simulation under single-phase ow
condition was conducted for one specic condition in a water departure from
nucleate boiling (DNB) test to examine the applicability of the CFD model for
predicting the CHF rod position. Energy ux around the rod surface in a water
DNB test is the sum of the intrinsic energy ux from a rod and the extrinsic
energy ux from other rods, and increments of the enthalpy and decrements of
ow velocity near the rod surface are assumed to aect CHF performance. CFD
makes it possible to model the complicated ow eld consisting of a spacer grid
and a rod bundle and evaluate the local velocity and enthalpy distribution around
the rod surface, which are assumed to determine the initial conditions for the two-
phase structure. The results of this study indicate that single-phase CFD can play
a signicant role in designing PWR spacer grids for improved CHF performance.
Toth and Aszodi (2008)[17] investigated the ow eld in subchannels of VVER-
440 pressurized water cooled reactors fuel assemblies (triangular lattice, P/D=
1.35). Impacts of the mesh resolution and turbulence model were studied in or-
der to obtain guidelines for CFD calculations of VVER-440 rod bundles. Results
were compared to measurement data published by Trupp and Azad in 1975. The
12
study pointed out that RANS method with BSL Reynolds stress model using a
sucient ne grid can provide an accurate prediction for the turbulence quan-
tities in this lattice. Applying the experiences of the sensitivity study thermal
hydraulic processes were investigated in VVER-440 rod bundle sections. Based
on the examinations the spacer grids have important eects on the cross ows,
axial velocity and outlet temperature distribution of subchannels therefore they
have to be modeled satisfactorily in CFD calculations.
Cheng et. al. (2009)[19] had carried out experimental investigations per-
formed on thermal-hydraulic behavior in tight hexagonal 7-rod bundles under
both single-phase and two-phase conditions. Freon-12 was used as working uid
due to its convenient operating parameters. Tests were carried out under both
single-phase and two-phase ow conditions. Rod surface temperatures are mea-
sured at a xed axial elevation and in various circumferential positions. Test
data with dierent radial power distributions are analyzed. Measured surface
temperatures of unheated rods are used for the assessment of and comparison
with numerical codes. In addition, numerical simulation using sub-channel anal-
ysis code MATRA and the computational uid dynamics (CFD) code ANSYS-10
is carried out to understand the experimental data and to assess the validity of
these codes in the prediction of ow and heat transfer behaviour in tight rod
bundle geometries. Numerical results are compared with experimental data. A
good agreement between the measured temperatures on the unheated rod surface
and the CFD calculation is obtained. Both sub-channel analysis and CFD calcu-
lation indicates that the turbulent mixing in the tight rod bundle is signicantly
stronger than that computed with a well established correlation.
E. Merzari et. al. (2008)[20] had examined the ow in tight rod bundles char-
acterized by long-term, large-scale coherent patterns in the stream-wise direction.
The issue of simulating these structures through unsteady CFD simulations em-
ploying periodic boundary conditions in the stream-wise direction is be addressed.
The validity of the approach is assessed through the comparison of a large eddy
simulation (LES) for similar ow conditions inside a simplied geometry and ex-
perimental data. A powerful statistical tool (proper orthonormal decomposition)
is used to analyze the time varying solution. The ow eld has been decom-
posed into a series of normal modes, identifying the structures responsible for
the ow transfer between sub-channels. Additional insights on the physics of
13
these coherent structures are obtained. An unsteady Reynolds averaged Navier-
Stokes simulation (URANS) of the ow in a rod bundle has then been carried out.
The comparison between numerical results and experimental results [Krauss, T.,
Meyer, L., 1998.] proves that accuracy can be achieved for averaged statistics
such as stream-wise velocity, turbulent intensity and wall shear stresses.
Cheng and Tak (2006)[18] CFD analysis was carried out for thermal-hydraulic
behavior of heavy liquid metal ows, especially lead-bismuth eutectic, in sub-
channels of both triangular and square lattices. Eect of various parameters, e.g.
turbulence models and pitch-to-diameter ratio, on the thermal-hydraulic behavior
was investigated. Among the turbulence models selected, only the second order
closure turbulence models reproduce the secondary ow. For the entire parameter
range studied in this paper, the amplitude of the secondary ow is less than 1
percent of the mean ow. A strong anisotropic behavior of turbulence is observed.
The turbulence behavior is similar in both triangular and square lattices. The
average amplitude of the turbulent velocity uctuation across the gap is about
half of the shear velocity. It is only weakly dependent on Reynolds number
and pitch-to-diameter ratio. A strong circumferential non-uniformity of heat
transfer is observed in tight rod bundles, especially in square lattices. Related
to the overall average Nusselt number, CFD codes give similar results for both
triangular and square rod bundles. Comparison of the CFD results with bundle
test data in mercury indicates that the turbulent Prandtl number for HLM ows
in rod bundles is close to 1.0 at high Peclet number conditions, and increases
by decreasing Peclet number. Based on the present results, the SSG Reynolds
stress model with semi-ne mesh structures (y+ 15) is recommended for the
application of HLM ows in rod bundle geometries.
C.C. Liu and Y.M. Ferng (2009)[21] applied computational uid dynam-
ics(CFD) methodology in investigating the detailed knowledge of thermal-hydraulic
characteristics in the rod bundle, especially with the spacer grid. These localized
information, including ow, turbulence, and heat transfer characteristics, etc.,
can assist in the design and the improvement of rod bundles for nuclear power
plants. In this paper, at three-dimensional (3D) CFD model with the Reynolds
stresses turbulence model is proposed to simulate these characteristics within the
rod bundle and subsequently to investigate the eects of dierent types of grid on
the turbulent mixing and heat transfer enhancement. Two types of grid designs
14
are used herein, including the standard grid and split-vane pair one, respectively.
Based on the CFD simulations, the secondary ow can be reasonably captured
in the rod bundle with the grid. Thesplit-vane pair grid would enhance both
the ow mixing and the heat transfer capability more than the standard grid
does, as clearly shown in the simulation results. In addition, compared with the
results of experiment and correlation, the present predicted result for the Nus-
selt(Nu) number distribution downstream the grid shows reasonable agreement
for the standard grid design. However, there is discrepancy in the decay trend
of Nu number between the prediction and measurement for the split-vane pair
gird. This would be improved by adopting the ner mesh (y+ 1) simulation
and Low-Reynolds form turbulence model.
Cesna (2010)[22] had presented the procedure of the cellular calculation of
thermo hydraulic parameters of a single phase gas ow in a fuel rod assembly.
The procedure is implemented in the DARS program. The program is intended
for calculation of the distribution of the gaseous coolant parameters and wall
temperatures in case of arbitrary, geometrically specied, arrangement of the
rods in fuel assembly and in case of arbitrary, functionally specied in space,
heat release in the rods. In mathematical model the ow cross-section of the
channel of intricate shape is conventionally divided to elementary cells formed by
straight lines, which connect the centers of rods. Within the limits of a single
cell the coolant parameters and the temperature of the corresponding part of the
rod surface are assumed constant. The entire fuel assembly is viewed as a system
of parallel interconnected channels. Program DARS is illustrated by calculation
of a temperature mode of 85-rod assembly with spacers of wire wrapping on the
rods.
2.4 Comments on the work done so far
The thermal hydraulic characteristic of can be signicantly dierent from that
of conventional circular pipe ows. The experimental study has been made from
past a few decades and turbulence mixing phenomena was tried to understand.
However, many researchers have tried to check the applicability of computational
methods to simulate the mixing phenomena. CFD code has been used only in
past few years for detailed study of thermal-hydraulic characteristics in fuel rod
15
bundles. From literature reviewed, it was found that only a few work [authors]
has been done in computational aspects in understanding the crossow behav-
ior of uid inside the fuel rod and have concluded that only by including the
anisotropy into the model is able to predict the crossow turbulent behavior.
But on the other hand, the observations made by some others like Toth and As-
zodi, Liu et. al., Cheng et. al.,Cheng and Tak, Kazuo Ikeda et. al. ([17] [21]
[18] [?] [16]) etc. show that the CFD code is able to predict the cross ow but
in context of the computational time and cost it is very expensive. Hence, it is
computationally viable to estimate the turbulence mixing coecient by CFD
simulating experimental conditions and use 1-D code subchannel analysis code
like COBRA in estimating the temperature and CHF margin in each subchannel.
In view of these observations, it appears that capability of CFD code to simulate
the cross ow behavior and estimating the coecient is not well established yet
and more studies need to be carried out in this direction.
2.5 Closure
This chapter details the literature survey and motivation for this project work.
16
Chapter 3
COBRA CODE and
GOVERNING EQUATIONS OF
CFD
This chapter gives breif details about COBRA code, equation it solves, relevance
of cross ow and importance of turbulence mixing coecient . In the next sec-
tion governing equations of CFD and breif introduction to turbulence is detailed.
3.1 Introduction to COBRA
COBRA is one of the popular sub-channel analysis code used for determining the
enthalpy and ow distribution in rod bundles for both steady state and transient
conditions. COBRA-IV-I is an interim version of COBRA for thermal-hydraulic
analysis of rod bundle nuclear fuel elements and cores prepared for the Energy Re-
search and Development Administration and for the Nuclear Regulatory Commis-
sion. This code is developed by C.L.Wheeler, C.W.Stewart, R.J.Cena, D.S.Rowe,
A.M.Sutey and D.S.Trent of Battelle Pacic Northwest Laboratories, Richland,
Washington. The foundations of the COBRA computer program are the conser-
vation equations used to describe uid and energy transport and the numerical
methods employed to solve them. The conservation of mass, linear momentum,
and energy are described by balance laws written for an Eulerian control volume.
the governing equations, with appropriate boundary conditions and constitutive
relationships, are then put into a framework for numerical solution. To pro-
vide the greatest possible computational capability and maintain the ecient use
of computing facilities, the COBRA-IV code contains two numerical methods.
17
One is an implicit technique similar to that used in COBRA-IIIC; the other, an
explicit method, is new. Both algorithms have an important function and, in gen-
eral, the implicit scheme is used for steady-state and slow transient simulation.
The explicit method is used for transients where ows are expected to reverse
direction.
3.1.1 Basic Equations
The basic equations of the mathematical model are derived by applying the gen-
eral equations of continuity, energy, and momentum to a subchannel control vol-
ume. These equations are derived by Rowe for subchannels and are also applied
in the present treatment. The equations are:
Continuity
A
h
t
+
m
x
+ [DC]
T
w = 0 (3.1)
Energy
A
h
t
+
mh
x
+ [DC]
T
h

w = q

(3.2)
AxialMomentum
m
t
+
mu
x
+ [DC]
T
u

w + A
p
x
= F (3.3)
TransverseMomentum
w
t
+
u

w
x
+
v
y
w
y
[DC]P = c (3.4)
In the above equations [DC] is a matrix operator which performs the lateral
nite dierence operation, [DC]
T
is the transpose of [DC] and performs a sum-
ming operation istead of dierencing, F is the axial friction and gravity force, c
is the lateral friction force, and q is the heat transfer from all sources.
The diversion crossow between adjacent subchannel per unit length, w is
important factor in these equations. To evaluate w there is no direct method and
is approximated mentioned in the next section.
18
3.1.2 Single-Phase Turbulent Mixing
Several forms of equations for specifying the turbulent cross ow are included.
The presently available forms in COBRA for calculating w include:
w

k
= s
k
G (3.5)
w

k
= aRe
b
s
k
G (3.6)
w

k
= aRe
b
D.G (3.7)
w

k
= aRe
b
s
k
z
k
D.G (3.8)
where
Re =
D.G

(3.9)
D = 4
A
i(k)
+ A
j(k)
Pw
i(k)
+ Pw
j(k)
(3.10)
G =
m
i(k)
+ m
j(k)
A
i(k)
+ A
j(k)
(3.11)
=
1
2
_

i(k)
+
j(k)

(3.12)
And a and b are input constants. Since a denitive mixing correlation does
not exist and other forms are available, the user should set up correlations of his
choice. Also included in the subcooled mixing is the thermal conduction. When
it is included, the conduction coecient is given by
19
c
k
=
k
i(k)
+ k
j(k)
2
s
k
z
k
K
g
(3.13)
Where K
g
is a geometric correction factor. Note that the distance z
k
is used
in both equations 3.8 and 3.13. This is the centroid-to-centroid distance between
subchannels. Care should be taken to select this value for its intended use. For
example, z
k
could be selected as the eective mixing distance.
3.2 Introduction to CFD and turbulence
In engineering application turbulent ows are common. In transportation of liq-
uids or gases with pumps, compressors, pipelines the ows are generally turbulent.
An important characteristic of turbulent is its ability to transport and mix uid
much more eectively than do laminar ows. Most important in turbulence is
the separation between small scale turbulence and large scale turbulence motions.
The large scale motions are strongly inuenced by the geometry of the ow on
the other hand small scale motions behavior is determined entirely by the rate at
which they receive energy from the large scales.
3.2.1 Reynolds Average Navier Stokes Simulation (RANS)
One important feature of the turbulent ow is: velocity, pressure etc. at a point
varies with time. In most of the applications the uctuating quantities (time
varying components) are not of much practical interest. Only average quantities
of velocity, pressure, temperature, etc are needed. With this rationale, the scalar
variables like velocity components, pressure etc., are split into mean and uctuat-
ing components and substituted into the Navier-Stokes equations. The governing
equations for the mean ow components are obtained by taking the Reynolds (en-
semble) average over the Navier-Stokes equations. The equations thus obtained
are called Reynolds-Averaged Navier-Stokes (RANS) equations. The Reynolds
averaging can be thought of as time-averaging is done over a period which is large
compared to the time period of random uctuations associated with turbulence
but small with respect to that of the associated unsteady mean ow. This aver-
aging results in the creation of unknown quantities u

i
u

j
involving product of
turbulent velocity uctuations. The quantities are called Reynolds stresses.
20
Thus in RANS approach only the mean ow components are resolved while
the eect of the uctuating components on the mean ow components is modelled
using some algebraic relations. In RANS approach, there are two major categories
in which the unknown Reynolds Stress term that appears in RANS equations is
modeled. Those are namely Eddy Viscosity Models (EVM) and Reynolds-Stress
Models (RSM).
3.2.2 Eddy-Viscosity Models (EVM)
Nowadays EVMs are most widely used to model the Reynolds Stress because
of its simplicity and lower computational requirements. In EVMs the Reynolds
stress/turbulent stress is modeled using the eddy viscosity hypothesis which cor-
relates the Reynolds stress with the mean shear strain rate through a parameter
of proportionality known as Eddy Viscosity or Turbulent Viscosity. This assump-
tion is called the Boussinesq hypothesis. Mathematically it can be written as,
u

i
u

j
=
t
_
u
i
x
j
+
u
j
x
i

2
3
u
k
x
k

ij
_

2
3
k
ij
where
t
is called as turbulent/eddy viscosity and k is the turbulent kinetic energy
dened as,
k =
u

i
u

i
2
There is a sub classication among these models as Isotropic and Anisotropic
Eddy Viscosity models. In the former one the eddy viscosity is postulated to
be isotropic as in the k , k , SST turbulent models, whereas in the later
the anisotropy of eddy viscosity is taken into account, as in the non-linear k
turbulence model [12] and Shih et. al. [23].
Another kind of classication in EVMs is based on the number of partial dif-
ferential equations (PDEs) to be solved for the turbulence quantities like algebraic
or zero-equation models, one-equation models, two-equation models etc.
Algebraic or zero-equation models
One-equation models
Two-equation models
21
In algebraic/zero-equation models no PDE is solved, instead the turbulent
viscosity is directly specied using an algebraic expression, e.g., mixing length
models which use Prandtls mixing length hypothesis.
In one-equation models, one PDE for a turbulent quantity is solved, eg., the
Spalart-Almaras model in which one PDE is solved directly for the Eddy Viscosity.
In some models, the equation for turbulent kinetic energy is solved which provides
the velocity scale and the length scale is prescribed directly using the Prandtls
mixing length hypothesis.
In two equation eddy viscosity turbulence models (EVM), the eddy viscosity is
modelled using two transport equations, one each for a turbulence quantity den-
ing velocity and time scales. Most EVMs are based on a transport equation for
the turbulent kinetic energy (k) which forms the velocity scale and an additional
transport equation for a second turbulence quantity, from which length/time
scales can be obtained, e.g., dissipation rate of turbulent kinetic energy in k
model, specic dissipation rate of turbulent kinetic energy in k model etc.
3.2.3 Reynolds-Stress Models (RSM)
In RSM approach each of the six Reynolds stress components has one transport
equation. So, there is no modelling at the level of second order moments. But
the third order correlations appearing in those transport equations again have to
be modelled. These models are therefore referred to as second-moment closures.
RSM equations are numerically more sti and generally more dicult to converge
as compared to EVMs.
3.3 Closure
The present chapter has presented the introduction of COBRA, the basic equa-
tions solved by it, the denition of as dened in COBRA and its relevance.
Introduction to CFD and turbulence equations(Reynolds Average Navier Stokes
Equation) is presented.
22
Chapter 4
ANALYSIS AND VALIDATION
WITH THE EXPERIMENTS
This chapter is devoted to the illustration of methodology, application of method-
ology, analysis and validation with the experimental results.
4.1 Introduction to Analysis
Many experimental have been carried out to investigate the turbulent mixing
phenomenon. To ascertain the methodology of estimating the turbulent mixing
coecient by CFD code the following cases have been considered and compared
(i) Triangular lattice bare fuel pin in isothermal case, Roidt et. al.[1], (ii) Square
lattice with egg crate spacer grids with non-uniform heating, Castellana[2].
4.2 Description of Methodology
4.2.1 Grid Preparation
The yplus value recommended to capture the secondary vortices developed in the
small narrow gaps is approx. 1, this is also in agreement with the literature, Liu
et.al. (2009)[21]. It is also important to have a aspect ratio of the cell approx.
1. The spacing in the axial direction need to be compact to capture global ow
pulsation phenomana.
23
Results of various grid conguration
The table below 4.2.1 details the average obtained by considering various grid
sizes. The phenomena of turbulence mixing is specic to the grid size. This is in
good agreement with the literature, Cheng and Tak (2006) [18], Toth and Aszodi
(2009) [17].
The geometry was resolved with four dierent meshes (refer table 4.2.1) in
order to study the inuence of the grid resolution. The cross sectional resolution
of the meshes was dierent while the height of the axial layers was the same in
all cases. The height of the axial layers is less important in this case because
fully developed ow was modeled solving the Reynolds-Averaged NavierStokes
(RANS) equations.
Mesh Type Size (cells in each face) Average
Very Coarse (Mesh 1) 93 6e-5
Coarse (Mesh 2) 453 4e-4
Fine (Mesh 3) 2745 3e-3
Very Fine (Mesh 4) 11341 3.1e-3
Table 4.1: Results of various grid conguration to assesss the optimize the mesh
size
4.2.2 Choice of Turbulence Model
The ow inside the closely packed fuel lattice geometry shows strong anistropy
behavior. The eddy viscosity model used to simulate the ow must consider
this anistropic factor to capture secondary ows which are driven by turbulence.
Various isotropic eddy viscosity model has been used and found the turbulence
mixing coecient obtained to be a order less than the value predicted the correla-
tions given in the literature. The basic model which encompasses the anisotropic
behavior is Reynolds Stress Model (RSM) (please refer to page 22). RSM is a
seven equation (7-eqn) model which do not use the Blassius approximation to
equate the scalar turbulence viscosity like any other LEVM (Linear Eddy Vis-
cosity Model). Infact, it is recommended that secondary or even tertiary closure
model will be able to predict these secondary phenomena to much more satis-
faction. However, in this thesis the calculations are based on RSM model and
secondary and higher closer model are out the scope of this work.
24
Results of various isotropic model
Based on the dimensionless turbulence mixing quantities it can be stated that
the accuracy of the numerical results are strongly aected by the resolution of
the applied mesh. The largest dierences between the measured and calculated
values can be observed in the case of the coarsest mesh (Mesh 1). Using ner
grids (Mesh 2, Mesh 3, Mesh 4) discretization errors decrease and for this reason
the discrepancies between measured and calculated values decrease, too (refer
table 4.2.2). The Mesh 2 is chosen for further analysis so as to optimize the
computational time and accuracy.
Turbulence Model Average
k 8e-4
k standard 9e-5
k RNG 7e-4
SST 9e-4
RSM 3e-3
Table 4.2: Results of various Turbulence model to obtain
4.2.3 Plotting vs. Axial length
Subchannel interfacing plane 0-1, 0-2 and 0-3 is extracted as shown in the gure
4.1. The values of velocities u, v and w , corresponding area of each cell is ob-
tained in the plane as shown in the gure 4.2. The normal velocity across the
plane is calculated by taking components of the velocities u and v. The absolute
value of this normal velocity V
n
is multiplied by corresponding area of the cell.
Area averaged normal velocity for each row of the cells in the plane is obtained.
This value is normalized by the average axial velocity to yield value. This is
plotted for each row of the plane axially as shown in the gure 4.3.
25
Figure 4.1: Subchannel connection
Figure 4.2: Cell conguration in the plane extracted.
Figure 4.3: A Typical Plot showing vs. Axial length of the plane
26
4.3 Case (i): Triangular Lattice - Isothermal
Case, Roidt et. al.[1]
4.3.1 Description of Experimental set-up
The experiment was performed on a large scale air ow model of a liquid metal
cooled fast breeder reactor rod bundle. A cross section of the facility is shown in
Fig. 4.4 and the axial arrangement in Fig. 4.5. The test section contained 39 2.5
inch diameter rods on triangular pitch at nominally 3.14 inch centers. The rows
were held in place at the downstream end by two hexagonal grids as indicated in
Figure 4.5. Air was drawn through the test section at velocities of up to 100 ft/sec
to yield overall duct Reynolds numbers on the order of 70,000. The numbered
rods in 4.4 could be slid in and out of the test section from the upstream end and
each was equipped with a 0.0625 inch diameter pitot-static probe which could be
moved radially from the rod surface. The tracer gas used in the experiment was
ethane (C
2
H
4
) with a molecular weight of 30.07 which is felt to be close enough
to that of air so that bouyancy should be negligible. It was injected into the
subchannel numbered (0) in 4.4 through an eight inch long 0.0625 inch diameter
tube positioned axially 0.56 inches from the rod directly above number 3 rod in
the same gure. The ethane samples were collected through the pitot static probe
dynamic head orices at the same mesh positions as the velocity readings were
taken.
Tracer gas concentrations were to be read at an axial location where velocity
data were known and both entrance eects and upstream grid eects were neg-
ligible. Such a location was 27 inches upstream of the grid. Using a pitot-static
probe as a sampler centered in subchannel (0) at this axial location, the axial
location of the injector, its position within the subchannel (0) and the tracer gas
ow rate were varied until a high concentration was obtained at the subchannel
(0) centerline with negligible concentrations at the outside clearance gaps of
subchannels (1), (2), anc (3). High concentrations and gradients were desirable
only at the clearance gaps between the (0) subchannel and those immediately ad-
jacent to simplify the data reduction procedure. The nal injector position was
68.3 inches upstream of the sampling probe with the injector position rotated
27
approximately 15 clockwise from the subchannel (0). centerline to the position.
A pressure correction on the ethane ow rate of approximately 3.45 liters/minute
was not obtained since only axial changes in concentrations within the test sec-
tion entered into the calculations. To obtain axial variation in the above mean
values the sampling probes had to be moved downstream to a position where such
a variation was detectable but where the values of both the concentration and
the concentration gradients at the clearance gaps did not change suciently to
prevent using mean values (the average of the upstream and downstream values)
in Equation 3.1. The distance to the downstream measuring plane, x, was taken
as six inches where the readings were repeated to yield the results summarized
in Fig. 4.3.3 and illustrated in Fig. 4.8 and Fig. 4.9.
Figure 4.4: Cross section of the test section of the experimental set-up
28
Figure 4.5: Schematic Layout of the Duct in the experimental set-up.
4.3.2 Description of Model and CFD input parameters
Cross sectional view of the model under consideration is shown in the Fig. 4.6.
Isometric view of the model is shown in the Fig. 4.7.
Figure 4.6: Cross sectional view of the model case(i)
29
Figure 4.7: Isometric view of the model case(i)
Parameter Value
Cmu 0.09
C1-Epsilon 1.44
C2-Epsilon 1.92
C1-PS 1.8
C2-PS 0.6
C1-PS 0.5
C2-PS 0.3
TKE Prandtl Number 1
TDR Prandtl Number 1.3
Table 4.3: The coecients of RSM turbulence model
The model is 2.6m long axially and containes 2682118 elements. The turbu-
lence model chosen is RSM (Reynolds Stress Model). The cocients used in the
RSM is shown in the table 4.3.2.
Input Parameters of the CFD is tabulated in the table 4.3.2 below:
4.3.3 Results and comparison with the experiment
The obtained from the experiment at upstream and downstream position is
tabulated in the table 4.3.3
30
Geometrical Details Flow Details
Fuel pin dia. : 2.5 in Inlet velocity: 100 ft/sec
Pitch : 3.14 in Working uid: air at 20 deg. Celsius (isothermal case)
Length of the test section: 225 ft Reynolds number: 70000 (based on hydraulic dia.)
Hydraulic dia. : 1.24 in
Table 4.4: Input Parameters for the case(i)[1]
Upstream position
Clearance gap number W

ij
(lb/ft-sec)
0-1 .1696 .0063
0-2 .0944 .0039
0-3 .0815 .0031
Downstream position
0-1 .1861 .0069
0-2 .0938 .0038
0-3 .0876 .0034
Table 4.5: Results of obtained from the experimental set-up [1]
Ethane concentration in the downstream position is plot in the Fig. 4.8
Ethane concentration in the upstream position is plot in the Fig. 4.9
The Concentration gradient across the clearance gap is plot in the Fig. 4.10
Normal Velocity through the axial length across the clearance gap is plot in
the Fig. 4.11
31
Figure 4.8: Ethane concentration in the downstream position [1]
Figure 4.9: Ethane concentration in the upstream position [1]
32
Figure 4.10: The Concentration gradient across the clearance gap [1]
Figure 4.11: Normal Velocity through the axial length across the clearance gap
Axial velocity in the downstream position is plot in the Fig. ??
Axial velocity in the upstream position is plot in the Fig. ??
33
Plot of Vs. Axial Length is shown in the Fig. 4.12
Figure 4.12: Plot of Vs. Axial Length for case (i)
4.3.4 Grid sensitivity of the model
The various grid conguration ranging from very coarse mesh to very ne mesh
has been studied to check the mesh sensitivity of the model. (Fig. 4.13, Fig.
4.14, Fig. 4.15 and Fig. 4.16)
34
Grid Type Size(cells on each face) Representative Value
Mesh 1 874 9e-4
Mesh 2 1390 3e-4
Mesh 3 2944 5.1e-3
Mesh 4 49640 5.1e-3
Table 4.6: vs. various grid conguration of case(i)
Figure 4.13: Cross section of the very coarse mesh (Mesh 1) conguration of
case(i) model
4.3.5 Conclusions
The value obtained by the methodology proposed in this work is matching with
experimental work.
Comparison between the result of obtained by CFD and experimental work
is tabulated in the Table 4.3.5 below:
35
Figure 4.14: Cross section of the coarse mesh (Mesh 2) conguration of case(i)
model
Figure 4.15: Cross section of the ne mesh (Mesh 3)conguration of case(i) model
4.4 Case (ii): Square Lattice with egg crate spacer
grids (four Nos.) with Non-uniform Heat-
ing
4.4.1 Description of Experimental set-up
A cross-section of the test geometry with all pertinent physical dimensions is
shown in g. 4.17. The simulated fuel rods were fabricated from type-347 stain-
36
Figure 4.16: Cross section of the very ne mesh (Mesh 4) conguration of case(i)
model
S. No. Subchannel Interaction (estimated by CFD) (by experiment [1])
1 0-1 5.2e-3 6.3e-3
2 0-2 5.1e-3 3.9e-3
3 0-3 3.1e-3 3.1e-3
Table 4.7: Comparison between the result of obtained by CFD and experimental
work
less steel tubing internally supported by closely tting ceramic cylinders, and
were heated directly by the lengthwise passage of electric current. The indicated
10:1 transverse power prole was aected with tubing of two dierent wall thick-
nesses. Rod spacing was maintained by four simple egg crate spacers designed
to minimize ow disturbance and located at 15 in. intervals beginning 12 in.
from the upstream end of the assembly.
The temperature at the exit of each of the 36 subchannels was measured by a
0.070 in. dia. Calibrated platinum resistance thermometer, accurate to -/+0.30F,
located in the center of the channel by a thermometer support grid. All temper-
atures were recorded automatically using a multichannel sequential scanner with
digital voltmeter and printer. Experiments were conducted in the high pressure
ow loop of the Columbia University Heat Transfer Research Facility (Fig. 4.18).
This system is nominally rated at 2200 psia and can provide ow rates up to 400
37
gal/min and 3.5 MW of electric power (since uprated to 12 MW).
The procedure adopted here is to measure subchannel temperature and then
attempt to predict the experimental measurements with a subchannel code by
varying coecients in an assumed mixing correlation.
Figure 4.17: A cross-section of the test geometry
Figure 4.18: Schematic sketch of the experimental test set-up
38
4.4.2 Description of Model and CFD input parameters
Cross sectional view of the model under consideration is shown in the Fig. 4.19.
Figure 4.19: Cross sectional view of the model case(ii)
A isometric view of the model is shown in the Fig. 4.20.
The model is 1.8 m long axially and containes 2621125 elements. As per
the experiment 4(four) egg crate spacers of height 10 mm is incorporated. The
spacers position is as per experiment. The turbulence model chosen is RSM
(Reynolds Stress Model). The cocients used in the RSM is shown in the table
4.3.2.
Input Parameters of the CFD is tabulated in the table 4.4.2 below:
Input Parameters of the CFD is tabulated in the table 4.4.2
39
Figure 4.20: A isometric view of the model case(ii)
Geometrical Details Flow Details
Fuel pin dia. : 10.72 mm Inlet velocity: 3.704 m/sec
Pitch : 14.3 mm Working uid: water at 438 deg. Kelvin
Length of the test section: 1.818 mm at inlet and vaying properties (C
p
,,k,)
as a function of temperature (438 to 573 deg. Kelvin
Heat ux(power 1x): 621303 W/m
2
Heat ux(power 10x): 6213030 W/m
2
Hydraulic dia. : 10.57 mm Reynolds number: 248000 (based on hydraulic dia.
Table 4.8: Input Parameters for the case(ii) [2]
4.4.3 Results and Discussion
An overall measure of t was determined by contrasting the experimental enthalpy
dierences with the dierences predicted by the COBRA code at a particular
constant value of , by means of the following relationship,
8

i=0
8

j=i
[(h
e(i)
h
e(j)
) (h
p(i)
h
p(j)
)]
2
(4.1)
In the above equation 4.1, h
e(i)
and h
e(j)
are experimental enthalpies in sub-
channels i and j, and h
p(i)
and h
p(j)
are the corresponding values predicted by the
COBRA code. The summation is eected over all possible pairs from the eight
unit cells. A value of S equal to zero represents a perfect t of all the data. Rel-
ative rather than absolute enthalpy dierences wesre used in an eort to negate
40
the eects of peripheral heat loss.
Plot of vs. error s in the Fig. 4.21
Figure 4.21: Plot of vs. error s
Plot of vs. rows of the intersecting plane for a bare model in the Fig. 4.22
Plot of vs. row number of the intersecting plane in the Fig. 4.23
Plot of vs. row number of the intersecting plane in the Fig. 4.24
From the above 4.24 the repeating itself after it encounters spacer each time
can be clearly seen.
Plot of vs. row number of the intersecting plane in the Fig. 4.25
41
Figure 4.22: Plot of vs. rows of the intersecting plane for a bare model
Figure 4.23: Plot of vs. rows of the intersecting plane for the model with
spacers
In the above g. 4.25 the values of
1
and
3
are higher than
2
and
4
. This
is due to the reason that the dierential heating of 10:1 power prole exists across
42
Figure 4.24: Plot of vs. row number of the intersecting plane showing details
near rst and second spacer
Figure 4.25: Plot of vs. row number of the intersecting plane showing details
near rst spacer
the horizontal direction 4.17.
43
4.4.4 Grid sensitivity of the model
bare model
The bare model is checked for mesh sensitivity by considering four grid congu-
ration as tabulated in the table below (4.4.4:
Various grid congurations are shown in the Fig. 4.26, Fig. 4.27, Fig. 4.28
and Fig. 4.29.
Grid Type Size(cells on each face) Representative Value
Mesh 1 93 2e-6
Mesh 2 453 4e-5
Mesh 3 2745 8e-4
Mesh 4 11341 9e-4
Table 4.9: vs. various grid conguration of case(ii) bare model
Figure 4.26: Cross section of the very coarse mesh (Mesh 1) conguration of
case(ii) bare model
44
Figure 4.27: Cross section of the coarse mesh (Mesh 2) conguration of case(ii)
bare model
Figure 4.28: Cross section of the ne mesh (Mesh 3)conguration of case(ii) bare
model
Model with spacers case(ii)
The various grid conguration ranging from very coarse mesh to very ne mesh
has been studied to check the mesh sensitivity of the model with spacer, case (ii).
(Fig. 4.30, Fig. 4.31, Fig. 4.32 and Fig. 4.33)
45
Figure 4.29: Cross section of the very ne mesh (Mesh 4) conguration of case(ii)
bare model
Grid Type Size(cells on each face) Representative Value
Mesh 1 336 2e-5
Mesh 2 570 2e-4
Mesh 3 1722 8e-3
Mesh 4 13892 8e-3
Table 4.10: vs. various grid conguration of case(ii)
4.4.5 Conclusions
The value with spacer grid is almost ten times the value obtained for the bare
model 4.22. The eect of spacer can be felt downstream the spacer grid. The
value slowly builts up as it approaches the spacer. Just before the spacer the
46
Figure 4.30: Cross section of the very coarse mesh (Mesh 1) conguration of
case(ii) model
Figure 4.31: Cross section of the coarse mesh (Mesh 2) conguration of case(ii)
model
drops to a value as close as to zero and then suddenly shoots up to a very high
value to the order of ten times. Downstream to the spacer the value drops and
saturates to a almost constant value ( 0.008). This value of =0.008 can be taken
for analysis as a input for the COBRA code and is almost same as obtained from
47
Figure 4.32: Cross section of the ne mesh (Mesh 3)conguration of case(ii) model
Figure 4.33: Cross section of the very ne mesh (Mesh 4) conguration of case(ii)
model
the experiment (0.0077). From the g. 4.24 it is very evident that the prole of
repeats itself when it encounters a spacer. It is of interest to know if result
is same when only one spacer can be used and length of the test section can be
reduced so that it is more economical with respect to computation.
48
4.5 Representative Models
4.5.1 Square Lattice - Isothermal Case
Description of Model and CFD input parameters
The model is 1 metre long and has a egg crate spacer grid at 400 mm. from the
inlet. The spacer grid is 10 mm. long and is 0.5 mm. thick.
Cross sectional view of the model under consideration is shown in the Fig.
4.19.
A isometric view of the model is shown in the Fig. 4.20.
Input Parameters of the CFD is tabulated in the table 4.5.1
Geometrical Details Flow Details
Fuel pin dia. : 10.72 mm Inlet velocity: 3.704 m/sec
Pitch : 14.3 mm Working uid: water at 438 deg. Kelvin (isothermal)
Length of the test section: 1.818 mm Reynolds number: 248000 (based on hydraulic dia.
Hydraulic dia. : 10.57 mm
Table 4.11: Input Parameters of the CFD for the isothermal case of representative
model
Results and Discussion
Plot of vs. Axial Length of the intersecting plane with one spacer grid is shown
in the Fig. 4.34.
Grid sensitivity of the model
The various grid conguration ranging from very coarse mesh to very ne mesh
has been studied to check the mesh sensitivity of the model.
49
Figure 4.34: Plot of vs. Axial Length of the intersecting plane with one spacer
grid - isothermal case
Grid Type Size(cells on each face) Representative Value
Mesh 1 992 5e-4
Mesh 2 1811 1e-4
Mesh 3 3162 8e-3
Mesh 4 45845 8e-3
Table 4.12: vs. various grid conguration of representative model (isothermal)
Conclusions
The representative value for can be taken as 0.008. This value is in agreement
with the actual model 4.23 and experiment. Hence, It is recommended that a
representative model can be taken with only one spacer grid to obtain the value
which can be used for the COBRA code for subchannel analysis. The length from
input to the spacer grid is suggested to be 100 times the hydraulic dia. whereas in
this model it is 400 times. The exit length from the spacer grid is recommended
to be 200 times and in this case it is 800 times.
50
4.5.2 Square Lattice - Uniform heating Case
Description of Model and CFD input parameters
The model and mesh is same as isothermal case. Refer Fig. 4.19 and Fig. 4.20
Input Parameters of the CFD is tabulated in the table 4.5.2
Geometrical Details Flow Details
Fuel pin dia. : 10.72 mm Inlet velocity: 3.704 m/sec
Pitch : 14.3 mm Working uid: water at 438 deg. Kelvin
Length of the test section: 1.818 mm at inlet and vaying properties (C
p
,,k,)
as a function of temperature (438 to 573 deg. Kelvin
Heat ux(power): 621303 W/m
2
Hydraulic dia. : 10.57 mm Reynolds number: 248000 (based on hydraulic dia.
Table 4.13: Input Parameters of the CFD for the uniform heating case of repre-
sentative model
Results and Discussion
Plot of vs. Axial Length of the intersecting plane with one spacer grid is shown
in the 4.35.
Grid sensitivity of the model
Refer to the table 4.5.1.
Conclusions
The value for all the subchannel interfaces is seen to be in agreement within
a very thin margin. The representative 4.35 is in agreement with the actual
model 4.23, isothermal case 4.34 and experimental data.
4.5.3 Square Lattice - Non-uniform heating Case
Description of Model and CFD input parameters
The model and mesh is same as isothermal case uniform heating case. Refer Fig.
4.19 and Fig. 4.20
51
Figure 4.35: Plot of vs. Axial Length of the intersecting plane with one spacer
grid -uniform heating
Input Parameters of the CFD is tabulated in the table 4.5.3
Geometrical Details Flow Details
Fuel pin dia. : 10.72 mm Inlet velocity: 3.704 m/sec
Pitch : 14.3 mm Working uid: water at 438 deg. Kelvin
Length of the test section: 1.818 mm at inlet and vaying properties (C
p
,,k,)
as a function of temperature (438 to 573 deg. Kelvin
Heat ux(power 1x): 621303 W/m
2
Heat ux(power 10x): 6213030 W/m
2
Hydraulic dia. : 10.57 mm Reynolds number: 248000 (based on hydraulic dia.
Table 4.14: Input Parameters of the CFD for the non-uniform heating case of
representative model
Results and Discussion
Plot of vs. Axial Length of the intersecting plane with one spacer grid is shown
in the Fig. 4.36.
52
Figure 4.36: Plot of vs. Axial Length of the intersecting plane with one spacer
grid - non-uniform heating
Comparison of values of isothermal case, uniform heating, non-uniform heat-
ing and bare model is shown in the Fig. 4.37.
Figure 4.37: Comparison of values of isothermal case, uniform heating, non-
uniform heating and bare model
53
Grid sensitivity of the model
Refer to the table 4.5.1.
Conclusions
The representative than can be taken from isothermal case, uniform heating
case and non-uniform heating case is not very signicantly dierent. Hence, it is
evident and can be concluded that isothermal analysis is sucient to obtain the
representative value for input into COBRA code.
54
Chapter 5
PREDICTION OF
TURBULENCE MIXING
BEHAVIOUR USING CFD FOR
SPACER LOCATION IN A
TYPICAL PWR
This chapter contains the application of the methodology developed in the thesis
onto a typical PWR whose case is under study.
5.1 Description of problem under study
The geometrical and operational data is given below for the problem under study:
Fuel pin diameter: 9.1 mm
pitch: 12.75 mm
Fuel assembly: Hexagonal channel, 37 pin cluster
Channel length: 2000 mm
Spacer type: petal type spacer grids (Nos. 10, interval of 200 mm)
spacer thickness: 0.25 mm
spacer height: 10 mm
Inlet velocity: 5 m/s
Inlet temperature: 273
0
Celsius
Operating pressure: 157 bar(a)
Power: 3120 MW(th)
A cross section of the channel with the spacer grid is shown in the gure 5.1
55
Figure 5.1: Cross section of the channel with the spacer grid
5.2 Description of CFD models
5.2.1 Bare model
A bare model of the given geometry is taken. The cross section of the model
under consideration is shown in the gure 5.2
Figure 5.2: Bare model of the problem under study
56
5.2.2 With one spacer
There are three kinds of subchannels possible for the given geometry. All the
subchannel conguration is chosen to estimate the turbulent mixing coecinet
. The representative models of the given geometry is taken for analysis. The
cross section of the model under consideration is shown in the Fig. 5.3 and Fig.
??csptye2) and its isometric view in the Fig. 5.5 and 5.6. The representative
model chosen is based on the conclusions in the page numbers 50 and 54 of
chapter 4.
Figure 5.3: Cross sectional view of the model type-1
5.3 Results and Discussion
The is plot against cells of the row in the axial length for both the models(Fig.
5.7 and 5.8). The plot here is very much similar to the plots which are shown in
57
Figure 5.4: Cross sectional view of the model type-2
Figure 5.5: Isometric view of the model type-1
the chapter 4. The representative value of =0.002 can be taken for analysis as a
input for the COBRA code. In the model type-2 the subchannel 0-3 shows more
lateral ow since the gap is wider for that particular subchannel (Fig. 5.4). The
value of for this geometry, spacer type and ow conguration is not available
in the literature. For the sake of comparison to the nearness to the solution this
value can be compared with the correlation obtained by Roger and Tahir (1975).
This comparison can be made because the p/d ratio used in the experiment is 0.4
which is same here in this problem under study, however the ow conditions and
working uid is very dierent(air in the case of experiment and water in this case)
The value by the correlation underpredicts (0.0008) the mixing coecient, .
58
Figure 5.6: Isometric view of the model type-2
Figure 5.7: The plot againt row numbers along axial direction for model type-1
From this it is very evident that this correlation value is misleading for the
cases which are out of range of experimental conditions. To verify this subchannel
analysis need to be carried out with both values of (obtained in this work and
59
Figure 5.8: The plot againt row numbers along axial direction for model type-2
by correlation given by Roger and Tahir) in COBRA code to plot the temperature
distribution at the exit.
Secondary ow vortices in innite rod bundles have been predicted by a num-
ber of theoretical analyses. Nevertheless experimental verication was dicult,
since the magnitude of the secondary velocities appeared to be less than the accu-
racy of the experimental techniques used. To capture these phenomenon in CFD
was notorously dicult. Only indications of the maximum velocity magnitude
have been available but no report on successful direct measurement is known to
the author by CFD.
Fig. 5.9 shows the vortex at the inlet of the spacer. The inuence of spacer
on the ow pattern can be clearly seen. At the exit the vortex is much more
dominant and is shown in the gure 5.10. The helical movement of the coolant
is dominantly seen at the downstream of the spacer. This shows the magnitude
of mixing that is occuring in the subchannel space. The complex ow pattern
can be visualized through this Fig. 5.11. The ow changes direction and creates
strong vortices and secondary ows due to strong anisotropy in the ow pattern
which is shown in the Fig. 5.13
60
Figure 5.9: Vortex at the entry of spacer wall
Figure 5.10: Vortex at the exit of spacer wall
61
Figure 5.11: Helical movement of the coolant
Figure 5.12: Vortex downstream the spacer
62
Figure 5.13: The ow changing direction downstream the spacer
5.4 Grid sensitivity of the model
5.4.1 Model Type-1
Mesh sensitivity analysis is carried out by considering various sizes of grid given
in the table below.
Various grid congurations are shown in the Fig. 5.14, Fig. 5.15, Fig. 5.16
and Fig. 5.17.
Grid Type Size(cells on each face) Representative Value
Mesh 1 250 9e-4
Mesh 2 311 5e-4
Mesh 3 912 2e-3
Mesh 4 1520 2e-3
Table 5.1: vs. various grid conguration of model type-1
63
Figure 5.14: Cross section of the very coarse mesh (Mesh 1) conguration of
model type-1
Figure 5.15: Cross section of the coarse mesh (Mesh 2) conguration of model
type-1
64
Figure 5.16: Cross section of the ne mesh (Mesh 3)conguration of model type-1
Figure 5.17: Cross section of the very ne mesh (Mesh 4) conguration of model
type-1
5.4.2 Model Type-2
The various grid conguration ranging from very coarse mesh to very ne mesh
has been studied to check the mesh sensitivity of the model. (Fig. 5.18, Fig.
5.19, Fig. 5.20 and Fig. 5.21)
65
Grid Type Size(cells on each face) Representative Value
Mesh 1 320 6e-4
Mesh 2 1020 3e-4
Mesh 3 4156 2e-3
Mesh 4 14514 2e-3
Table 5.2: vs. various grid conguration of model type-2
Figure 5.18: Cross section of the very coarse mesh (Mesh 1) conguration of
model type-2
5.5 Conclusions
The representative value obtained for the problem under study is 2e-3. Vortex
formation, secondary ow and helical movement of the coolant is plot. The
magnitude of these secondary ows is in agreement with the literature [7].
66
Figure 5.19: Cross section of the coarse mesh (Mesh 2) conguration of model
type-2
Figure 5.20: Cross section of the ne mesh (Mesh 3)conguration of model type-2
5.6 Closure
In this chapter the value is obtained for a typical PWR which is the problem
under study by the methodology developed in this thesis work.
67
Figure 5.21: Cross section of the very ne mesh (Mesh 4) conguration of model
type-2
68
Chapter 6
SUBCHANNEL ANALYSIS OF
A TYPICAL PWR USING THE
SUBCHANNEL CODE
COBRA-IV
This chapter contains description of subchannel analysis carried out on a typical
PWR under study. The analysis is carried out with (i) With value of from the
literature and (ii) value predicted in this thesis The results of COBRA-IV code
case (ii) is also compared with CFD analysis.
6.1 Subchannel Analysis
6.1.1 Description of input le
One twelth symmetric model of the channel is taken for the subchannel analysis.
To show the eect of coolant mixing due to the presence of spacer subchannel
analysis is done on both bare channel and spacer grids for comparison (refer g.
6.2 and 6.1). To compare the results of subchannel analysis and CFD results a
one twelth symmetric model of 215 mm long and with one spacer grid is taken
(refer next section 75). Uniform heat ux is taken from both in radial and axial
direction of the fuel pins. Other factors like axial momentum factor, lateral
resistance factor etc. is taken as default value of 0.5.
69
6.1.2 Results and Discussion
With value of from the literature
For the geometrical conguration and ow characteristics there is no correlation
for turbulence mixing readily available in the literature. Since the p/d ratio of
this geometry is 0.4 and there is correlation available by Rogers and Tahir (1975)
for triangular conguration of the same p/d ratio, for comparison this value can
be taken for analysis. The obtained from the formula 6.1 is 0.0023.
=
w

.
D
h
S
.Re
0.1
(6.1)
w

= 0.005(
S
d
)
0.106
Re
0.9
(6.2)
where d is fuel rod dia. S is clearance gap D
h
is hydraulic dia. Re is Reynolds
number
Subchannel No. Average temperature at exit
(degree Farhenheit)
1 644.91
2 646.18
3 645.15
4 643.07
5 640.5
6 640.87
7 634.4
8 632.71
9 633.42
10 632.39
Table 6.1: Average temperature at the exit and CHF margin of 1/12th symmetric
model with ten spacer grids by COBRA analysis with value of beta given in the
literature
With value predicted in this thesis
The value obtained in this thesis is 0.002 (refer to g. 5.7 and 5.8 in page
number 59).
70
Subchannel No. Average temperature at exit
degree Farenheit
1 644.7
2 646.5
3 645.25
4 643.17
5 640.57
6 640.75
7 633.98
8 632.22
9 633.03
10 631.99
Table 6.2: Average temperature at the exit and CHF margin of 1/12th symmetric
model with ten spacer grids by COBRA analysis with beta obtained from CFD
The results for the bare model is tabulated below in the table 6.1.2.
Subchannel No. Average temperature at exit
(degree Farenheit)
1 675.23
2 673.12
3 671.51
4 674.24
5 676.32
6 677.86
7 670.41
8 671.18
9 668.84
10 669.24
Table 6.3: Average temperature at the exit and CHF margin of 1/12th symmetric
bare model by COBRA analysis with beta obtained from CFD
6.1.3 Conclusions
The temperature distribution at the exit of the channel and margin on CHF in
the channel is plot for the value of given in the literature and obtained by
the results of CFD analysis. The temperature distribution is not very sensitive
71
to value and dierence in maximum temperature is found to be 40
o
Farenheit
with spacers and without spacerss.
6.2 CFD Analysis
CFD codes are capable of handling subchannel analysis. Many researchers have
been successful in predicting temperature distribution in the subchannels and
CHF margins [17] [21] [18]. In this section an attempt is made to estimate
temperature prole at the exit by CFD and is compared with the results of
COBRA-IV code. However, the complete channel is not modelled but only a
representative section length is chosen to make it economical in computation.
Objective of this analysis is to prove that chosing a representative value from
CFD and using in the subchannel analysis is more computationaly viable than
detailed analysis in CFD to estimate temperature proles.
6.2.1 Description of CFD model
A representative model is chosen for analysis. The channel is hexagonal (refer
Fig. 5.1 and as seen in the gure 1/12th symmetric model can be chosen for
analysis. The model is 215 mm. long for both bare model and model with one
spacer. A total of 1814528 cells in bare model and 2816234 in model with one
spacer are present.
A cross sectional view of the bare model is shown in the Fig. 6.1.
A cross sectional view of the model with one spacer is shown in the Fig. 6.2.
A isometric view of the model with one spacer is shown in the Fig. 6.3.
72
Figure 6.1: A cross sectional view of the bare model of 1/12th symmetric channel
for analysis
Figure 6.2: A cross sectional view of the model with one spacer of 1/12th sym-
metric channel for analysis
6.2.2 Results and Comparison with the COBRA-IV re-
sults
Typical run time for the analysis is 200 hours for residual error of 1e-5.
The results of CFD analysis is tabulated (refer table 6.2.2) giving the exit
temperature and results those of subchannel analysis.
73
Figure 6.3: A isometric view of the model with one spacer of 1/12th symmetric
channel for analysis
6.2.3 Grid sensitivity of the model
Mesh sensitivity analysis is carried out by considering various sizes of grid given
in the table below (refer Fig. 6.2.3).
Various grid congurations are shown in the Fig. 6.4, Fig. 6.5, Fig. 6.6 and
Fig. 6.7.
6.3 Conclusions
The channel length considered here is only representative. A detailed analysis of
all the subchannels in the channel and with full actual length one would invite
a very high hardware requirement and a huge amount time. On the other hand
by carrying out a isothermal analysis with a representative model and estimating
74
Subchannel No. Avg. temp. at exit (CFD) Avg. temp. at exit (COBRA)
(degree Farenheit) (degree Farenheit)
1 642 644.7
2 644 646.5
3 645 645.25
4 642 643.17
5 638 640.57
6 637 640.75
7 630 633.98
8 629 632.22
9 631 633.03
10 630 631.99
Table 6.4: Comparison table showing average temperature at the exit given by
CFD analysis and COBRA analysis
Grid Type Size(cells on face) Average temperature at exit
Mesh 1 670 500
Mesh 2 1018 510
Mesh 3 4624 630
Mesh 4 12142 632
Table 6.5: Average temperature at the exit vs. various grid conguration of
1/12th symmetric model with one spacer for analysis
75
Figure 6.4: Cross section of the very coarse mesh (Mesh 1) conguration of 1/12th
symmetric model with one spacer for analysis
Figure 6.5: Cross section of the coarse mesh (Mesh 2) conguration of 1/12th
symmetric model with one spacer for analysis
and doing a analysis with subchannel analysis code is very computationally
viable.
6.4 Closure
In this chapter the details of subchannel analysis is given. The subchannel anal-
ysis is carried with (i) default value of and (ii) with value obtained by CFD
analysis in the 5. The detailed CFD analysis is carried out for 1/12th symmetry
76
Figure 6.6: Cross section of the ne mesh (Mesh 3) conguration of 1/12th sym-
metric model with one spacer for analysis
Figure 6.7: Cross section of the very ne mesh (Mesh 4) conguration of 1/12th
symmetric model with one spacer for analysis
model and the results are compared with COBRA code results for verication.
77
Chapter 7
CONCLUSIONS AND SCOPE
FOR THE FUTURE WORK
7.1 Conclusions
This thesis focuses on the numerical simulation of turbulence ow phenomena
in tightly packed fuel pin subassemblies of a typical PWR and turbulence mix-
ing coecient has been estimated for spacer locations. A methodology has been
suggested to estimate this turbulent mixing coecient. The methodology before
applying to the actual problem under study it has been extensively checked for its
applicability on following cases: (i) isothermal case of triangular lattice bare fuel
pins (ii) non-uniform heating case of square lattice fuel pins with egg crate spacer
grids. These cases are simulated by Computational Fluid Dynamics (CFD). The
CFD methods employed here are those of anisotropic eddy viscosity models of
turbulence and anisotropic Reynolds-Averaged Navier-Stokes (RANS) equations
approach. Complicated turbulent ow structure is due to strong anisotropy in
the non-uniform channel geometry that is characterized by wide open channels
connected by a narrow gap. The secondary ows in sub-channels play an im-
portant role in transporting small eddies generated in the wider region toward
the narrow gap. The methodology is applied on to the problem under study.
The estimated by this methodology is input into the COBRA code to obtain
the desired results like subchannel temperature, CHF margin and other outputs.
These results are compared with the results predicted by CFD code to check the
reliability of the evolved.
78
7.2 Scope of the Future Work
The methodology can be applied on various P/D ratios, various Reynolds number
for various spacer grid types and a correlation can be developed which can be
generalized. The turbulence model used in the CFD code is basic anisotropic
model (RSM). A more specic turbulence model like second order k and cubic
k can be used to study this turbulence mixing behavior.
79
References
80
References
[1] Roidt et al., Experimental determination of turbulent exchange coe-
cient in a model reactor rod bundle,Heat Transfer and Fluid Dynam-
ics Department, Westinghouse Research Laboratories, Pittsburgh, PA,
1970
[2] Castellana, SINGLE-PHASE SUBCHANNEL MIXING IN A SIMU-
LATED NUCLEAR FUEL ASSEMBLY,Nuclear Engineering and De-
sign Vol. 26, pp. 242-249, March, 1974.
[3] W.J. SEALE, TURBULENT DIFFUSION OF HEAT BETWEEN
CONNECTED FLOW PASSAGES, Nuclear Engineering and Design
Vol. 54, pp. 183-195, 1979.
[4] A.C. TRUPP and R.S. AZAD, THE STRUCTURE OF TURBULENT
FLOW IN TRIANGULAR ARRAY ROD BUNDLES, Nuclear Engi-
neering and Design Vol. 32, pp. 47-84, 1975.
[5] T. Krauss and L. Meyer, Experimental investigation of turbulent trans-
port of momentumand energy in a heated rod bundle, Nuclear Engi-
neering and Design Vol. 180, pp. 180-206, 1988.
[6] Creer et al., TURBULENT FLOW IN A MODEL NUCLEAR FUEL
ROD BUNDLE CONTAINING PARTIAL FLOW BLOCKAGES, Nu-
clear Engineering and Design Vol. 52, pp. 15-33, 1979.
[7] V. VONKA, MEASUREMENT OF SECONDARY FLOW VOR-
TICES IN A ROD BUNDLE, Nuclear Engineering and Design Vol.
106, pp. 191-207, 1988.
81
[8] Elvis E. Dominguez-Ontiveros and Yassin A. Hassan, Non-intrusive
experimental investigation of ow behavior inside a 55 rod bundle with
spacer grids using PIV and MIR, Nuclear Engineering and Design Vol.
239, pp. 888-898, 2009.
[9] Hartmut RAMM et al., SINGLE PHASE TRANSPORT WITHIN
BARE ROD ARRAYS AT LAMINAR TRANSITION AND TURBU-
LENT FLOW CONDITIONS, Nuclear Engineering and Design Vol.
30, pp. 186-204, 1974.
[10] C.W. RAPLEY and A.D. GOSMAN, THE PREDICTION OF FULLY
DEVELOPED AXIAL TURBULENT FLOW IN ROD BUNDLES,
Nuclear Engineering and Design Vol. 97, pp. 313-325, 1986.
[11] A.K. Mohanty and K.M. Sahoo, TURBULENT FLOW AND HEAT
TRANSFER IN ROD-BUNDLE SUBCHANNELS, Nuclear Engineer-
ing and Design Vol. 106, pp. 327-344, 1987.
[12] Kye Bock Lee and Ho Cheol Jang, A numerical prediction on the tur-
bulent ow in closely spaced bare rod arrays by a nonlinear k- model,
Nuclear Engineering and Design Vol. 172, pp. 351-357, 1997.
[13] Michio Sadatomi et al., Prediction of the single-phase turbulent mix-
ing rate between two parallel subchannels using a subchannel geometry
factor, Nuclear Engineering and Design Vol. 162, pp. 245-256, 1996.
[14] Hae-Yong Jeong et al., A correlation for single phase turbulent mix-
ing for square rod arrays under highly turbulent conditions, Nuclear
Engineering and Design Vol. 162, pp. 305-317, 1996.
[15] E. Baglietto and H. Ninokata, A turbulence model study for simulating
ow inside tight lattice rod bundles, Nuclear Engineering and Design
Vol. 235, pp. 773784, 2005.
[16] Kazuo Ikeda et al., Single-phase CFD applicability for estimating uid
hot-spot locations in a 55 fuel rod bundle, Nuclear Engineering and
Design Vol. 236, pp. 11491154, 2006.
82
[17] Toth and Aszodi, CFD analysis of ow eld in a triangular rod bundle,
Nuclear Engineering and Design Vol. 240, pp. 352363, 2010.
[18] Cheng et al., CFD analysis of thermalhydraulic behavior of heavy liq-
uid metals in sub-channels, Nuclear Engineering and Design Vol. 236,
pp. 18741885, 2006.
[19] Shih-Kuei CHENG and Neil E. TODREAS, HYDRODYNAMIC
MODELS AND CORRELATIONS FOR BARE AND WIRE-
WRAPPED HEXAGONAL ROD BUNDLES - BUNDLE FRICTION
FACTORS, SUBCHANNEL FRICTION FACTORS AND MIXING
PARAMETERS, Nuclear Engineering and Design Vol. 92, pp. 227-
251, 1986.
[20] E. Merzari et al., Numerical simulation of ows in tight-lattice fuel bun-
dles, Nuclear Engineering and Design Vol. 238, pp. 1703-1719, 2008.
[21] C.C. Liu et al., Numerically simulating thethermalhydraulic charac-
teristics within the fuel rod bundle using CFDmethodology, Nuclear
Engineering and Design (article in the press).
[22] B. Cesna, Analytical model for calculation of the thermo hydraulic
parameters in a fuel rod assembly, Nuclear Engineering and Design
Vol. 240, pp. 37083715, 2010.
[23] Tsan-Hsing Shih et al., A new Reynolds stress algebraic equation
model, Comput. Methods Appl. Mech. Engrg. vol. 125, pp.287-302,
1995.
[24] Patankar S.V., Numerical Heat Transfer and Fluid Flow, Taylor and
Francis, 2004.
[25] Pope Stephen B., Turbulent Flows, Cambridge University Press, 2000.
[26] Wilcox D.C. and Chambers T.L., Critical Examination of Two-
Equation Turbulence Closure Models for Boundary Layers, AIAA
Journal, Vol. 15, No. 6, pp. 821-828, 1977.
83
[27] Wilcox D.C., Turbulence Modeling for CFD, DCW Industries Inc.,
1998.
[28] Patankar S.V. , Numerical Heat Transfer and Fluid Flow , Taylor
and Francis , 2004.
[29] Spallart P.R. and Allmaras S.R., A One-Equation Turbulence Model
for Aerodynamic Flows, AIAA-92-0439, 30th Aerospace Sciences Meet-
ing & Exhibit, Reno, January, 1992.
[30] Speziale C.G., On Nonlinear k l and k models of turbulence,
Journal of Fluid Mechanics, Vol. 178, pp. 459-475, 1987.
[31] X. Cheng and Y.Q. Yua, Local thermalhydraulic behaviour in tight 7-
rod bundles, Nuclear Engineering and Design, Vol. 239, pp. 19441955,
2009.
[32] Frank S. CASTELLANA and Joseph E. CASTERLINE, SUBCHAN-
NEL FLOW AND ENTHALPY DISTRIBUTIONS AT THE EXIT OF
A TYPICAL NUCLEAR FUEL CORE GEOMETRY, Nuclear Engi-
neering and Design, Vol. 22, pp. 3-18, 1972.
[33] D.C. Groeneveld et al., The 2006 CHF look-up table, Nuclear Engi-
neering and Design, Vol. 237, pp. 19091922, 2007.
[34] NEIL E. TODREAS et al., COOLANT MIXING IN LMFBR ROD
BUNDLES AND OUTLET PLENUM MIXING TRANSIENTS, Re-
port No. DOE/ET/37240 - 109 FR, August 1984.
[35] K. REHME and G. TRIPPE, PRESSURE DROP AND VELOCITY
DISTRIBUTION IN ROD BUNDLES WITH SPACER GRIDS, Nu-
clear Engineering and Design, Vol. 62, pp. 349-359, 1980.
[36] Seizo HIRAO and Noboru NAKAO, DIANA - A FAST AND HIGH
CAPACITY COMPUTER CODE FOR INTERCHANNEL, Nuclear
Engineering and Design, Vol. 30, pp. 214-222, 1974.
84
[37] W. Zeggel and C. Monir, Prediction of natural mixing in tightly packed
seven-rod bundles ( P/D = 1.10), Nuclear Engineering and Design, Vol.
126, pp. 361-377, 1991.
[38] R. NIJSING et al., LATERAL TURBULENT DIFFUSION FOR
LONGITUDINAL FLOW IN A RECTANGULAR CHANNEL, Nu-
clear Engineering and Design, Vol. 32, pp. 221-238, 1975.
[39] R. NIJSING et al., STUDIES ON FLUID MIXING BETWEEN SUB-
CHANNELS IN A BUNDLE OF FINNED TUBES , Nuclear Engi-
neering and Design, Vol. 5, pp. 229-253, 1967.
[40] W. EIFLER and R. NIJSING, EXPERIMENTAL INVESTIGATION
OF VELOCITY DISTRIBUTION AND FLOW RESISTANCE IN A
TRIANGULAR ARRAY OF PARALLEL RODS, Nuclear Engineering
and Design, Vol. 5, pp. 22-42, 1967.
85

Вам также может понравиться