You are on page 1of 77

TURBULENT INTERCHANGE

IN TRIANGULAR ARRAY ROD BUNDLES

+N

WALTER LEE KIRCHNER

B.S., United States Merchant Marine Academy

Ye

SUBMITTED IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE

DEGREE OF

MASTER OF SCIENCE

at the

MASSACHUSETTS INSTITUTE OF TECHNOLOGY

August, 1973

Signature of Author..... a =n
Signature redacted
DepartmentofNaclearEngineering
qo ~~ August 13, 1973

Cert.fied by...... uh a MR 4 Signature redacted


Thesis unaiigii
Accenprnd DV
Chairman, Departmental Committee
on Graduate Students

Archives
MASS. INST. TECH
SEP 18 1973
{Us RARIES
Title of Thesis: Turbulent Interchange
in Triangular Array Rod Bundles

Name of Author: Walter Lee Kirchner

Submitted to the Department of Nuclear Engineering, August 1973,

in partial fulfillment of the requirements for the Degree of Master

of Science.

ABSTRACT

Turbulent interchange is of current interest in nuclear


engineering for its importance in thermal-hydraulic performance.
The investigation of mixing in triangular array rod bundles reflects
the current utilization of this geometry in high power density
nuclear reactor cores.

A literature survey was conducted to assess the importance


of recent contributions of knowledge to the mixing field. An ex-
perimental program was established to provide a foundation for
further study of turbulent interchange in triangular array rod
bundles. A P/D = 1.05 test section and flow loop were constructed
and tested as the basis for this program.

Initial data collected reflected difficulties in the in-


strumentation systems, which are currently being corrected. Results
indicated that important areas to be investigated include: a systema-
tic sequence of P/D test sections, the importance of Reynolds number
variation, local concentration gradients for assessment of developing
flow effects, eddy diffusivities, and secondary flows.

Thesis Supervisor: Neil E. Todreas

Title: Associate Professor of Nuclear Engineering


#

ACKNOWLEDGMENTS

The author wishes to take this opportunity to thank

Professor Neil E. Todreas for his advice and assistance through-

out this study. The comments and advice of Professor Peter Griffith

often proved to be of great help and were much appreciated.

The use of equipment from the Engineering Projects Labora-

cory greatly expedited much of the construction work; to J.A. (Tiny)

Caloggero a special thanks.

The financial support of the Sloan Foundation which allowed

me to continue my studies is gratefully acknowledged. Thanks are

also due to the United States Atomic Energy Commission for their

financial support in procuring the equipment for this project.


TABLE OF CONTENTS

Title Page..veee.s 3 - ® 9 & }

AbStracte.eeeeees. i
“ 8=» S 4" a Oe ® ®& NN BB A ceaese2

Acknowledgments. ...... Cs 8 oe

Table of Contents... a A 5 8 9 % 4 8 % 6 5 © 8 P 0 @ 0 oo 8 2 a & a & 4 & 9 © © 8 © © © 0 © 8 °° ® 4

List of TAbleS eee eneeeeeeseoeeenens TENE » 6

List of FiguUresS..eeeeeiveeneseen » & V8 5 @ & & @& =» a 2 & & & ® ® 9 © © 8B © © © o > 7

Symbols.....- % 7 8% & )F 8 © bP» & BA & » &a 2 a 2 1 8B 8 9 6 ® & 8 ° 8 0 8 Q

I. IntroducCtion.eeeeeeeeseeeeses, nn a . . in
.10

[I. Theory

A. The Energy Equation... rm oa ease eeeesdld

B. Turbulent Interchange........-.. 16

IIT. Literature Review

A. Survey Articles... = 9 8 9 L021

B. Recent Experimental Studies and Correlations........ 21

C. Low Reynolds Number Mixing Phenomena.......... » - 27

IV. Experimental Apparatus and Procedure

A. Equipment Description........ » & 33

B. Experimental Procedure. .ceeeeeeeeene seen

J Data Analysis and Results

A. Data Evaluation Techniaqu- 45

B. Results....... *» 3} ®a 48
VI. Conclusion

A. Summary....... Cee cesseseeD2

B. Recommendations for Future Work.ieeeeeeeeeoeenennnen52

References..oeeeeee... x Pb ® @ ®™ ) zz & a - » & 2 a»a ) 8 1 &


LY

Appendices
A. Flow Meter Calibration.... con wwss37

B. Analysis Of Walton's Dat@...eeeeeerseceseeecceeennnseab3

C. Concentration Curve..... Ce. .68

D. Reynolds Number Conversion FactOTr...ue...eeeeeeeeeeesa?l

E. Mixing Data.........

F. Dye Injection Investigation.. -


tc eessalb
LIST OF TABLES

La Mixing ProcesseS.ssseceeesens a vowed

rz. Recent Experimental Investigations..... “ 2

ITI. Test Section Specifications........ - Bn a2 0»


, 37

B-I. Walton's Single Phase Water Dat@....cseeacee.vessasasrsaabb

B-ITI. Walton's Single Phase Air Data.. cesesenesD/

C-I. Concentration Curve Daté.ssssrsssnasssassasrsssssssvasns sd

E-T. Mixing Data.....- - ..75


LIST OF FIGURES

1. Triangular Array Rod Bundle (partial section)...........ll


2. Subchannel Energy BalanCe..e.ceeeesoasesascesosonesnssssld

}. Gap Laminarization........ Ch eee28

4. Secondary Flow... cviiviesessnerveseverennnss see eeneasdl8

3. Eddy Diffusivity Anisotropy. eeeeeeeseeuonnnaneseneseness29


5. Low Reynolds Number Phenomenon.. . 31

7 Experimental Loop Diagram (schematicC)....eeeeeeeeeeesa..34


3 Cross-section of Test SeCtiON.e.eeeeseeseecenesncaanneessl’

Tracer-Mass Balance.....- EY

10. Experimental ResUlES.ieee ene cots eeencaesnsessnnsenessasd

A-1. Port Flow Meter Calibration (9.5 GPM) .....evennren. .58

A-2. Starboard Flow Meter Calibration (9.5 GPM) ..vvsuveeeeeseb9

A-3. Port Flow Meter Calibration (6.0 GPM).....vvveereeeneee..b60

A-4. Starboard Flow Meter Calibration (6.0 GPM). ....evvevv... 61

A-5. Tracer Injection Flow Meter Calibration.. th eaeb2

C—l. Concentration CUT VE. .eeeeeeeeseeenseeseaneaeaneeesneenaaddl

D-1. Three SUbchannNels.c.eeeceeosssss tse es es sss eeesseescessesddl

F-1. Normalized Donor Subchannel Temperature Development.....77


-

SYMBOLS

A cross-sectional area

C - subchannel concentration

~ - subchannel exit concentration

Ce initial donor subchannel concentration


~,.
—- gap spacing

>o - specific heat (constant pressure)

D — diameter of fuel rod

Dy, — subchannel hydraulic diameter

be ~ friction factor

# —- gravitational constant

h - subchannel enthalpy
yA
- effective enthalpy

> - thermal conductivity

L - mixing length

Ly - effective mixing length scale

wm subchannel mass flow rate

pitch of fuel rod centers

~ pressure

on - Prandtl number

33 —

- heat
linear heat rate

rate per unit area

radial distance from fuel rod centerline


0

Re — Reynolds number

Sc — Schmidt number

St
T —- temperature

T(r) —- fluctuating temperature (radial variation)

n(n) — fluctuating velocity (radial variation)


~~
~ axial distance

J
1 —- distance between adjacent subchannel centroids

Zi - mixing distance

Fe
2 - effective mixing distance

— H
- thermal eddy diffusivity (radial component)

Fug - thermal eddy diffusivity (circumferential component)

£m - eddy diffusivity of momentum

~ - dynamic viscosity

- kinematic viscosity

5 - fluid density

od —- circumferential angle (zero at rod gap)

Wey — diversion cross flow

Ww—< — turbulent interchange

wey - effective turbulent interchange

Subscripts
- subchannel ¢

} - subchannel 4
10

[. INTRODUCTION

Turbulent interchange is of current interest in nuclear

engineering for its importance in thermal-hydraulic performance. Ac-

curate assessment of the degree to which this mixing phenomenon oc-

curs will benefit the designer in his thermal analysis of the re-

actor under steady state and transient conditions.

Although this phenomenon is generally applicable to power

reactors utilizing axial flow in parallel rod matrices, the specifi-

cation of triangular array reflects the recent trend to high power

density cores in fast breeder reactors. Fig. 1 illustrates a typi-

cal triangular array rod bundle, arbitrarily divided into subchannels.

Since most reactor cores are composed of an extremely

large number of fuel rods, the design and analysis computer codes

use a subchannel analysis. Digital computer programs such as

COBRA, HAMBO, HECTIC, HYDRA, SASS, and THINC calculate flow and en-

thalpy in the subchannels due to molecular and turbulent transport

of mass, momentum, and energy in the bundle. These codes utilize

mathematical expressions to take into account the various mixing

phenomena, as defined by Rogers and Todreas [1].

Natural mixing consists of those processes which occur

in the absence of flow perturbers such as spacer grids. Turbulent

interchange is the natural turbulent eddy transport between subchan-

nels, characterized by a zero net flow between the communicating

subchannels. In contrast, diversion cross-flow is a flow redis-


4

Figure 1

Triangular Array Rod Bundle

J
ys
\

HOC ) +

OO
*

\
\

DOC
1
\
\
\

a
) (OO)
~~
CHC J

DOC
NON
)/
/
£

Subchannels
1. Central
2. Edge
3. Corner
12
tribution process caused by radial pressure gradients between sub-

channels. This driving force can arise from a variety of sources

including : fuel rod bowing, variances in linear power generation

rate between fuel rods, or other flow area variations.

nn. 1
4iy €

Natural mixing effects Forced mixing effects

Non directional
flow effects Turbulent interchange Flow scattering

Directional
flow effects Diversion cross-flow Flow sweeping

Forced mixing effects are considered to be those mixing

processes caused by protuberances in the flow passage. Flow scat-

tering is a random process which may be effected by structural devices

such as grid spacers, which tend to increase the turbulent inter-

change over natural background values. Flow sweeping, resulting in

a preferred flow direction, is a directed process which may be as-

sociated with such devices as wire wrap spacers.

To obtain reliable results from these codes there is a

need for well defined expressions of the mixing processes. This

thesis is restricted to consideration of turbulent interchange. Al-

though it is evident that other factors such as position in the flow

field. Prandtl number, and flow field development should be accounted

for, to date, most correlations for turbulent interchange employ only

Reynolds number and geometry of the subchannels. The author makes

no pretensions for presenting a definitive correlation; however,


13
it is hoped that this experimental and analytical study adds know-

ledge to the mixing field.


14

[I. THEORY

A. The Energy Equation

A sample computer formulation of the energy equation will

be presented to indicate the importance of the mixing processes.

Following the notation of COBRA, as developed by Rowe [2], the energy

equation for the ith subchannel is:

Figure 2

wp he + &x meh de
| __ x+dy
EY
— w;. h. dx

wee hy dn -

(¢) (})
hid7
3 do”

fe
Pr
4

"a *
h .
15

‘hk: TR a / (1)
me; kh; + 2
on
mh dx + We h; dn + wish de = §dx + mh, tord hd
¥
where R is the effective enthalpy transported by diversion cross

flow across the gaps between subchannels. This quantity is normally

defined as

bh > 0
/
La}
a

(2)
| h; we<0d
The definition implies average or uniform values for h; or A, in

the subchannels. By definition, turbulent interchange incurs a zero

net flow, or

(43
Ww
f
= w ie

Using these results in Equation (1) yields

we: ’ ’ o

2 mh; 23 wy (h; - hh)- wih


¢

Note that the mixing processes have been treated in an additive

fashion; hence, they are considered to be independent of each other

in this analvsis.

Considering all channels adjacent to the ith, then

N
d mm:
dn
i—————.

>
4 1
4
(5° ’
16
Equation (4) now becomes

mY» dhe
¢ —= w— > Ww;&f (hn;J - h.). pe
> ww ;$ (h.~h
¢ ) (6)6
The term, 2 » is comprised of two parts: the heat transfer from the
/

fuel rod surfaces, ? , and the conduction heat transfer from one

subchannel to another via the gap between the fuel rods. For ex-

perimental studies using tracers, the above result may be recast

in terms of concentrations.

3 Turbulent Interchange

In order to use a subchannel analysis computer code the

user must supply a mathematical expression for the mixing process

considered. The usual procedure in the case of turbulent inter-

change is to perform an energy balance across the gap of the fuel

rods, equating the heat transport rate across the gap by the hypo-

thetical mixing rate to that of the transverse eddy fluctuation.

A suitable starting point in this derivation is a turbulent

neat transport rate equation such as developed in most texts dealing

with convective heat transfer [3]:

/ — —

34 (2) =
9}-_—_
<i (n) oo= k
A
dT
——
(~) -5 Pé
o nr, (K) T (x) (7)
P-D Jd
where the first term on the right-hand side is representative of

the conductive heat transfer and the second of the convective process.

The following assumptions are made for the derivation that follows:
1/

(1) steady state flow conditions exist, (2) the flow is incom-

pressible, and (3) the fluid properties are independent of temper-

ature.

The Boussinesque approach suggests using an eddy diffu~

sivity concept, where

wy (1) T(x) = —€,, (nL dT (A) 8°


Kk Yo
Note that temperature gradient on the right-hand side of Equation (8)

is a time averaged value. Using this result in Equation (7)

vield:.s-

3; (x) = gaon
(0) / kT
Lx ye(x) 7 § rE A I
)é (9)

The spatially averaged linear power generation rate at the gap

becomes

_ F/o

in-2fflrsEa)£2]a (10)

YY

F/2
24) --2epH 1 t Eu
ES —EW)L dTood(x)ho
dn (11)
/2
where the Prandtl number, Pr, is defined as

Pr=€Cp
1
1€
Following the notation of Todreas, et al. [4], an ef-
*
fective turbulent interchange, we » is defined to be the radially

averaged mixing flow rate for energy exchange across the gap, by

both turbulent and molecular effects. Equating the energy trans-


»
port by this hypothetical mixing, Ww; , to the result of Equation

(11) yields:

wi so (T:-T)= 954 1 7)

YY

wlA 2
(T:-T;)
(J+
o/s Fr
Eu)
v
1node a
bog

where 1; and 1 are the subchannel average temperatures.

The preceding results are exact within the framework of

the assumptions made; however, specialization is necessary to make

contact with prior work in the field. By assuming that

| Bh
XdLL 14-5= constant iq
. +)

and the eddy diffusivity is a constant, subchannel averaged value,

then Equation (13) becomes

Tw
we (P-D)[ 1
5 Hm) a
J) ¢ b:0 [ 2+ Eng
a—
| as)
T. -T
19
If the conduction term is neglected, Equation (15) may be simplified

to the "fundamental" equation of turbulent mixing (note that wg


refers only to turbulent mixing):

Eng
ra
ps

(P-D)¢ BR RA
(16:

where the effective mixing distance, Zz, . 1s

714 J
1 n
[ Ya]PU
It should be noted that the results predicted by Equations (15) and

(16) are dependent on the boundary conditions (i.e., tracer injection

technique) and will not necessarily predict the same answer for a

given geometry and flow condition.

The reader will notice that for a complicated geometry

such as a triangular array rod bundle, the assumptions employed

in obtaining this result greatly simplify the actual situation.

Eddy diffusivity in such a geometry has radial and circumferential

components, the relative magnitudes of which vary as functions of

position, Reynolds number, Prandtl number, and flow development.

Therefore, the use of a subchannel averaged value of eddy diffu-

sivity greatly reduces the complexity of the analysis (reference

to Fig. 5 indicates the anisotropy of the eddy diffusivity components

as a function of position).

The assumption of Equation (14) should be equally suspect,

since the gap region has a particularly complex flow field (see
20
Fig. 3). Finally, it should be noted that the effective mixing
*
distance, Zi » is just that, an effective value. It is not a

geometrically measurable quantity, though it may be compared to the

measured distance between subchannel centroids. This ratio may be


x
considered as a mixing length scale, Ls , defined by:

¥*

Ly4 = 24 (18
Y;;
Thus, further work in this field will require the ex-

perimenter to extend beyond considerations of geometry and Reynolds

number, only, to obtain a more definitive correlation for turbulent

interchange.
21

ITI. LITERATURE REVIEW

A. Survey Articles

For the reader's reference the author lists prior survey

work done in the field of mixing and deletes regurgitation of these

articles in favor of reviewing more recent works of allied interest.

Coates [5] published one of the first survey works in the mixing

field and was early to point out the problems of entrance effects

and developing flow fields on mixing. Moyer [6] implemented Elder's

[7] relationship for eddy diffusivity based on turbulent shear flow

in an open channel. This approach was subsequently employed by

many workers in the field to define eddy diffusivity. The Todreas

and Wilson [8] article, and the subsequent Rogers and Todreas [1]

review, provide a rather complete survey of work in the mixing field

through 1968.

R Recent Experimental Studies and Correlations

Petrunik [9] performed the first in a series of Canadian

experimental investigations on turbulent interchange. The test

section used for his tests consisted of two rectangular subchannels

divided by a slot of varying communicating lengths. Petrunik cor-

rectly pointed out the importance of entrance effects on observed

mixing rates. Shorter slot lengths yielded higher mixing rates,

a result that can be attributed to the wake caused by the opening


22
of the slot and the high concentration gradients of the developing

mixing field. By applying a systematic procedure of "subtracting"

entrance effects, Petrunik arrived at the following correlation for

single phase mixing:

J 13
St = 0.027Re (19)

where Re covered a range from 1 x 10% to 1.5 x 10°.

Walton [10] continued the work of Petrunik by extending

the investigation to two adjacent triangular array subchannels.

Although two different window lengths (57 and 115 hydraulic diameters)

were used in his P/D = 1.05 test section, no attempt was made to

remove entrance effects; as done by his predecessor, since apparently

Walton viewed them as negligible. This author disagrees, since

application of a subtracting procedure revealed that the values

reported for Re less than 5000 were often inflated (see Appendix B).

For Re greater than 5000, Walton correlated his data by St. Pierre's

[11] result:

St = 0.041 Re °°" (0)

Although Walton indicates that the flow region about Re = 5000 is

a transition zone, he erroneously concludes that molecular conduction

is not a significant factor in his results for flows below this

region.

Based on his earlier experimental work on mixing for

square~triangular and square-square two subchannel test sections,


23

Rowe [12,13] suggests the following correlation [14]:

5t = 2 0.0038Re™" 2 2)
Ly ei
where Ly is the relative centroid distance between subchannels,

normalized to a square-square array value of unity, and Dy, is the

hydraulic diameter of the square array subchannel. This correlation

is recommended for use in the Reynolds number range from 10

through 5 x 10° with bundles "mot too closely spaced." Rowe did

not use a window in his work, hence his correlation reflects the

inclusion of developing flow effects. He vaguely hints that a

secondary flow mechanism augments the measured turbulent inter-

chance.

Galbraith improved the simple two subchannel test section

technique in his work by using an "infinite square-square rod

bundle. Six rods were housed in a rectangular housing with the

capability of varying the clearance gaps. This arrangement allowed

communication with peripheral subchannels, resulting in a more

realistic flow field in the two subchannels of interest, without

compromising the precision of the areas. (Note that the peripheral

subchannels are not geometrically similar to the two subchannels

of interest, nor do they have equivalent hydraulic diameters.)

Significantly, Galbraith and Knudsen [16] point out that

for small gap spacings laminarization is occurring, impeding the nor-

mal mixing process by ''choling" the throat. Pressure imbalance

tests were performed to show the sensitivity of the mixing process


24
to small radial pressure gradients (fraction of an inch of water).

Diversion cross-flow was enchancedatlowerReynolds numbers,

10” to 10°, while at higher flows this effect is expectedly suppressed.

For a Reynolds number range of 10 to 5 x 10 in a square-square

array, they correlated their data (in the absence of pressure

imbalances) by:
—— - =

0. >:0
Ci 0.0113 Re (227 J

where the mixing distance is given by a relation of the form

= ca [4 exp {- fleRS £922 J

The work of Singh and St. Pierre [17] has added rein-

forcement to the contention that secondary flow mechanisms are

present in rod bundles, aiding the mixing process. Using a square-

square array, two subchannel test section, they investigated mixing

for air and water over a Reynolds number range of 2 x 10° to

3.8 x 10% for small P/D ratios (see Table II). The researchers

declined to present a correlation, deferring until more knowledge

of the structure of turbulence in the gap region was obtained.

The most recent general literature review is that of

Rogers and Rosehart [18], which includes an update of their corre-

lations in light of more recent data. The present correlation is

basically a modification of the Rogers and Tarasuk result [19]:


A

d= mm

Ag=—25tRe KD
Du; \ 3= ] Du 4
*/Z 2

[1 + (54)
Du: = Wa, D
where m = 0.2 and mm = 0.9. The value of mt. = 0.9 contrasts with

the earlier value of m.= (0.68. The correlation was updated to

cover more recent data by developing a relationship between the

mixing distance, zy , and the clearance gap to rod diameter

ratio, ce; /D , [20]:

rx

§ oo KT(5) (25° Pl

By modifying the results presented in a previous report [21], the

recommended correlations presented are:

for simple geometries, square-square.

- 0.894

A
0.005 (5)
D
(26 r

square-~-triangular

-1.36
No
-0.0026 (4)
D
(9"' )

and bundles,

-14¢6
Ne; = 0.0058 (4)
D
(28)
Table II. Recent Experimental Investigations

Researcher Geometry Experimental


I———————
Re Range Technique Comments (Data)

Petrunik [9] Simple square-square 10000- Considered entrance


0.80"x0.74" 150000
Tracer ,KNO (water)
Methane (air) effects (pp.55-59)
Walton [10] I'-T, two subchannels 2000-
P/D-1.05, D=0.776" 90000 Tracer,KNO, (water)
Methane (air)
Triangular array
(App.B herein)
Ingesson & Ring of rods in Heat Balance (air) Mixing factor
Hedburg [22] annulus correlated in
P/D=1.62,1.1 terms of geometry
D=0.47",0.79"
Van Der Ros & Simple square-square 5000- Heat Balance (water) 4 rod section
Bogaardt [23] J.394"x0.394" 30000 representative
of PWR
4 rods, S-S 76000-
P/D=1.33,D=0.394" 270000

Galbraith & 6 rods, S-S 8000- Tracer,Fluorescent Small gap spacings,


Knudsen [16] P/D=1.011,1.028, 30000 dve (water) pressure imbalance
1.063,1.127,1.228 tests
D=1.00" ([15],pp.144-148)
Singh & 1300-
St. Pierre [17]
S-S, two subchannels
P/D=1.02,1.04,1.10 38000
Tracer ,KNO, (water)
Methane (air)
Low flows and small
gap spacing
D=0.78",0.82",0.84" ([28],pp.92-99)
Rowe & Angle 5-S, S-T, 68000- Tracer, LiOH (water) ([13], C1-C9)
[12, 13] two subchannels 308000
P/D = 1.035, 1.15 : NN
p—

D = 0.564"
The authors recommend application of Equation (28) for

flows of Reynolds number greater than 2 x 10% and clearance-dia-

meter ratios of 0.08 to 0.4.

C. Low Reynolds Number Mixing Phenomena

f,aminarization in the gap region between subchannels can

occur at low flows or for small clearance gaps. The importance of

laminarization lies in the restricting, ''choking,'" effect that it

can have on the normal mixing process of turbulent eddy transport.

The extent of laminarization greatly affects the fuel and clad

temperatures. Fig. 3 illustrates a possible flow field, indicating

the turbulent, transition, and laminar regions.

As mentioned earlier, thermal eddy diffusivity appears to

be dependent on geometry, position, Prandtl number, and Reynolds

number. This situation is compounded by the anisotropy between

circumferential and radial diffusivities, as illustrated in Fig. _)

[24]. Since little work has been done to determine eddy dif-

fusivities of heat directly, correlations such as Dwyer's [25]


28
Figure 3

Gap Laminarization

\ 2~
"turbulent core

~ transition region

J
laminar sublayer

\
\

Figure 4

Secondary Flow

. |

; secondary flow field


|
|
"ha J

\ J

pO
20
Figure 5

Eddy Diffusivity Anisotropy

p——
30

40
5, P =30°
TN
SS——
A

P/D = 1.08
En,
\ Re = 50000
\
Pr = 1.0
0

v) 7 4a IJ 2

t=? 4 °

I'D 10 10°" 1.0

y=1
VoL
30

En_4_ 182 4 “
f2¢ J
Pr (En)VV /max

are utilized to relate thermal values to measured momentum diffusi-

vities. For a more detailed discussion of this topic the author

refers the reader to the work of Ramm and Johannsen [26].

Another phenomenon of importance to be considered in mixing

studies is secondary flow. It was first observed by Nikuradse [27]

for turbulent flows in straight channels of noncircular cross section.

Secondary flow is believed to be caused by variations in the local

wall shear stress gradient due to noncircularity of the flow channel.

Fig. 4 is a schematic representation of secondary flow in a triangular

rod array. The magnitude of the secondary flow is expected to in-

crease as the rod bundle spacing is decreased, but the degree of

penetration in the gap region will be related to the laminarization

present. The exact nature of this phenomenon merits further in-

vestigation.

With respect to Walton's work [10], application of an

entrance length effect removal process (as done by Petrunik [9]) will

show that the reported mixing for water at low flows is high.

Further analysis of this data yielded an interesting result, ex-

hibited as Fig. 6. At low flows (less than Re = 10000) the mixing

rates reported for air and water diverge. This can be explained by

the onset of laminarization in the gap region. Since the transport

process may proceed only by conduction in the laminar regime, the

larger mixing values for air reflect a lower Schmidt number.


31

Figure 6

Low Reynolds Number Phenomenon

1:
A JA

“A
p

"1
to

Fo »

ga’

0 water

A air

3
LC
. :

i 1 i.

Re
32
Rewriting Equation (15), including the conduction transport

term, offers an explanation.

-y = (2D
z%
5
Sc
, Exe
v
(30 r

Since the Schmidt number for methane in air is the order of unity,

while for a salt ion in water it is approximately one thousand, the

reader can observe the divergence as the turbulent eddy transport

decreases in significance. This result concurs with the conclusions

of Singh and St. Pierre [17], who observed a similar result for a

square—sdquare array.
33

IV. EXPERIMENTAL APPARATUS AND PROCEDURES

A. Equipment Description

Fig. 7 is a schematic diagram of the basic experimental

loop. The test section, as shown on the right-hand side of the

figure, was designed by the author. A good deal of effort was

expended in the design stage to incorporate those features of value

from previous researchers' models. From a simple two subchannel

model, through several transitions, the final design evolved, as

illustrated in Fig. 8.

The multi-channel test section represents an attempt to

model the actual flow field of a rod bundle as accurately as pos-

sible, without sacrificing dimensional tolerances or incurring

flow perturbations. This design philosophy was the rationale for

including the peripheral subchannels surrounding the two subchan-

nels of interest. For the case of the simple two subchannel model,

laminarization in the corner regions would distort the velocity

profile toward the mixing gap. It was also expected that the

asymmetry of the geometry would increase the magnitude of the

secondary flows. The net result indicated that a higher mixing

value would be obtained for the simple model versus the "infinite"

model (for a given subchannel flow rate).

The "infinite" model was not without its drawbacks, the

nost obvious being the increased complexity of the machining opera-


am

too
pot
|
BSB oO
~
PI
L

Eeg er
-
= oO
nn
_ ~ I}
/ B=) Lt

ON

[2] PH wn].

—— 7

J
5 de
~
I=
po

=
0
5 J
5
fo
tH
pow <
8 0
~ Oo
A

TT BS
No
bd

7 mixing
section 8
0

mt =
' _—

x
5
HA
(5
~~ LD
oS
:MD
a
DEVS
]
35
Figure 8

Cross-section of Test Section

Pitch-diameter ratio 1.35

Rod diameter 0 .750 1"

Hydraulic diameter 1.29 x 107% ft

O
SO ~ ~
O .
SSN
O
-—y

pa!

O ds
J
SA y
.

O)
~~
~
~
~ ’

<

SN Nf «

O
O hae
rt

2 "og lS— ra > Bp pana

LNASS
O DO O
36
tions. When this experimental investigation was initially conceived,

plans called for three, or possibly four, triangular array test

sections covering a range of pitch-diameter ratios and rod diameters.

Delays in machining of the first test section (P/D = 1.25, D = 0.250")

and failure of the resulting product to meet design specifications

sharply curtailed the scope of this investigation. The second test

section (P/D = 1.05, D = 0.750", illustrated in Fig. 8) benefited

from the experience gained in constructing the first. The problems

encountered included: (1) difficulty in maintaining satisfactory

tolerances due to the use of '"Plexiglas'" as the material of construc-

tion, (2) failure to have "fuel" rods centerless-ground in the first

test section resulted in an unacceptable tolerance deviation, and

(3) in retrospect, choice of a machine company lacking sufficient

expertise for jobs of this nature.

As mentioned previously, the test section was constructed

of "Plexiglas." The choice of this material for construction was

dictated by several factors: relatively low cost and ability to

visually observe processes occurring (i.e., visual dye or potential

for laser investigation). The relative ease in machining of

"Plexiglas" was negated by difficulties in maintinaing design toler-

ances. Important specifications of the test section are listed

below in Table III.

The test section was 4.5 ft. long. The flow field was

allowed to develop in an entrance length of 200 hydraulic diameters

after entering the test section from the entrance section. Inter-

changeable stainless steel windows of 100 and 50 hydraulic diameters

could be installed for observing the mixing processes. The choice


37

Table IIL

Test Section Specifications

-
Lo
Iriangular array (| J

Pitch~-diameter ratio 1.05

Pitch 0.7875"

Rod diameter 0.750"

Gap 0.0375"
-2
Hydraulic diameter 1.29 x 10 © ft

Length 4.5 ft

Entrance length (L/D) 200

Window lengths (L/Dy)


long 00

short 50

Window thickness 0.020"


38
of the longer window length was arbitrary, reflecting a balance

between cost factors (test section length) and the necessary length

for full development of the mixing field. The shorter window pro-

vided a means of roughly estimating entrance effects (see Literature

Review).

To provide for maintenance of proper clearances upon

disassembly and reassembly of the test section, three sets of

locating dowel pins were installed. After initial alignment in

the machine shop, this feature fixed the geometry for all suc-

ceeding reassemblies. Located four inches from either end of the

test section were two pairs of pressure taps for sensing differ-

ential pressures across the dividing window.

Identical entrance and exit sections provided for transi-

tion from loop piping to the test section. These units, also

fabricated of "Plexiglas," were five inches long. 3/4" NPT fit-

tings allowed coupling to loop piping at these interfaces. Two

1" diameter holes were bored in such a manner as to encompass the

exposed cross-sectional areas of each pair of three subchannels.

A 1/16" rubber gasket sealed the test section-entrance (or exit)

interface and prevented leakage between the separated flows. As

a further precaution to ensure proper matching of these units, the

connecting flanges were mounted in a stepped fashion to provide a

male— female joint between sections.

The water supply for the experimental tests consisted

of a 350 gallon aluminum tank fed by the local city water service.

Hot and cold water was available for temperature regulation as

desired. A Worthington D-820, close-coupled, centrifugal pump


39
(2" suction, 1" discharge, 6" impeller) supplied the fluid driving

energy. The prime mover was a General Electric 5 HP electric motor

(3600 RPM, 3 phase, 60 Hz, 220 volt).

The loop, as constructed by the author, was mounted on

plywood panels, with the necessary regulating devices centrally

mounted above the sampling sink. The flow split control centered

on two matching Fischer and Porter flow meters, F & P Model 10A3567

units, rated at 9.5 GPM each (these units could be derated to 6.0 GPM

with interchangeable floats, see Appendix A). Globe valves (1/2"

NPT brass) at the exit of each flow meter provided the primary

flow regulating capability. Similar globe valves downstream of

the test section provided additional flow control by adjusting the

level of back pressure on the system (see following section).

Of key concern in the loop design was the assurance that

the tracer would be completely mixed in the donor loop and at the

outlets of the test section. To achieve this objective, the tracer

was injected in either half of the loop, before the flow meter and

regulating valve, by a diffuser. The diffusers were 1/8" diameter,

closed-end, stainless steel tubes with multiple 1/32" holes drilled

about the periphery and length. The diffusers' axes were mounted

perpendicular to the flow to increase the degree of turbulent mixing

of the injected tracer with the main flow. Mixing screens were

installed in each half of the loop before and after the test section

to increase the degree of mixing of the tracer with its carrier.

The screens were fine mesh brass wire, fitted perpendicular to the

flow direction in the loop piping.


40
The tracer injection system utilized a 15 gallon galvanized

pressure tank as a reservoir. The tracer could be injected into either

half of the main loop by pressurizing the tank with air. Pressure

control was accomplished with a conventional regulator supplied by

service air at 100 psig. Flow control was established by a pre-

cision 1/8" needle-point metering valve downstream of a Fisher &

Porter, F & P Model 10A1017-A, "Tri-Flat" rotameter, rated at 200

cc/min (max). A 1/8" tantalum ball was used as a float to prevent

corrosive attack by the tracer.

Differential pressure regulation within the test section

was based on the use of inclined manometers. The manometers were

calibrated in hundredths of an inch of water, based on an inclina-

tion of 15° and an indicating fluid of 0.834 specific gravity.

Since the senseal fluid was water, it was necessary to use a manometer

fluid of specific gravity greater than one. A 2.96 specific gravity

standard fluid was used. This caused the sensitivity of the mano-

meter to be decreased roughly by one-fourth (sensitivity is directly

related to the ratio of specific gravities).

Sampling valves (1/8" brass needle-point) performed a

dual function of withdrawing samples and allowing fine control of

pressure imbalances in the test section. Two sample lines, each

connected to one half of the main loop after the flow meter and globe

valve, served to withdraw samples of the flow before it entered the

Lest section. Another matched pair served a similar purpose at the

discharge side of the test section (after the mixing screens).

The instrumentation system used for measuring the tracer

concentrations was based on a conductivity technique. The Leeds


41
and Northrup 4959 multi-purpose electrolytic conductance bridge

was the heart of this system. This instrument had a range from

0.5 to 105,000 micromhos with a limit of error of approximately

t0.5%7 in the range of interest (10™* mhos). The conductivity cell

employed was a Beckman CEL-VS1 unit with a cell constant of unity.

B. Experimental Procedure

Prior to any experimental runs, a concentration curve

was generated using known concentrations of sodium chloride in water

(see Appendix C). This established that the readings of the con-

ductance bridge varied linearly as the tracer concentration was

changed, within the range of interest. This was an important

result, for it greatly simplified the data analysis and evaluation.

Similar tests showed that the shape of the curve did not change with

changes in the background conductivity, but was simply displaced

to intercept the ordinate at the new background value. Had the curve

not been linear, it would have been necessary to convert all read-

ings to their corresponding concentration values instead of using

the readings directly (see Data Analysis and Results).

The sequential procedure used in taking experimental

data is outlined as follows:

Manometer Adjustment--The manometers were purged

completely of all entrained air bubbles by pressurizing

the system and venting to the atmosphere. By closing the

downstream globe valves (#3 and 4, numbers refer to

Fig. 7 nomenclature), and the pump discharge valve

(#5), the system could be hydrostatically pressurized


19
via the tracer injection system. Under this condition,

no differential pressures existed across the pressure

caps (except for minor variations in piping height),

allowing the adjustable sliding scales on the mano-

meters to be zeroed at the meniscus level of the mano-

meter fluid.

) Tracer Solution—--After depressurizing the system,

the desired tracer concentration was mixed for injection

in the tracer reservoir. The sodium chloride was

thoroughly mixed with an appropriate amount of water

before addition to the reservoir. This procedure

prevented variations in the injected concentration.

The salt was weighed on a standard laboratory balance

and mixed with water from the same supply that fed the

main storage tank.

d Loop Purge--After filling the main tank with water

at the desired temperature, the pump was started and

water was allowed to circulate through the loop.

All lines were completely purged of air bubbles at

this point, with special attention given to the

manometer lines.

4 Flow Adjustment--Primary flow control was accomplished

with the globe valves (#1 and 2) immediately down-

stream of the flow meters. Each flow was adjusted

to give equivalent flow meter readings at the selected

value. The globe valves (#3 and 4) downstream of the


L =
jo

test section were then throttled until the flow began

to just perceptibly fall below the desired value

selected on the flow meters. This procedure resulted

in a back-pressure on the system which damped out most

flow fluctuations and yielded a finer control of any

pressure differences at the test section.

Tracer Injection--The tracer reservoir was pressurized

10 psi above the pump discharge head. The donor

channel was selected and a regulated amount of tracer

was injected into it via the needle valve (#6) and

tracer flow meter.

La
“3 Pressure Balance--The isolation valves to the mano-

meters were now opened. Any gross pressure differ-

entials could be adjusted with the globe valves in the

main loop. Since the test section represented the

largest pressure drop in the loop, it had a net effect

of decoupling the fine control of the loop before the

test section from the remainder of the loop downstream

of the test section. The manometer before the mixing

window was very sensitive to the globe valves (#1 and 2)

at the flow meters, while the manometer at the exit

end was quite sensitive to the back-pressure (controlled

by globe valves #3 and 4). The sampling valves were

now ''cracked" open one-half turn. At this point it

was usually necessary to readjust these valves to

remove any pressure differences.


Ya
ol

/ Sample Withdrawal--When the system had reached steady

state under the conditions outlined above, samples

were taken at the sampling sink in 250 ml beakers.

These were immediately analyzed for conductivity and

the results were recorded. Of the four samples

withdrawn, the two upstream (of test section) readings

providedabackground value and a donor concentration

value. The two downstream samples gave the donor

channel outlet concentration, depleted by the mixing

process, and the concentration of the acceptor channel

after mixing.
o)
Alternate Injection--Under the same steady state

conditions outlined above, the tracer injection was

Immediately switched to the opposite channel. The same

procedure was employed, making certain that the flow

rates were identical with those of the preceding run.

This technique provided a means of checking for diversion

cross-flow, since similar results should be expected

(within the limits of experimental error) for in-

jection in either channel under identical flow conditons.


45

V. DATA ANALYSIS AND RESULTS

A Data Evaluation Technique

It is desirable to have an analytical technique that may

be quickly applied to evaluate data in the laboratory. The simple

two subchannel mixing equation provides such a result. If the test

section in this study is thought of as being comprised of two sub-

channels (there are actually six), then with certain qualifications,

the simple result may be used.

The following assumptions are necessarily made in using this

approach:
1 The mass flow rates in each channel are equivalent.

) Turbulent interchange is the only mixing mechanism

operable (diversion cross-flow is absent, molecular

conduction is considered negligible).

3. The tracer is thoroughly mixed with the subchannel

fluid, such that all eddy transport is assumed to have

an average concentration representative of the axial

location of interchange.

4. The tracer effect on fluid properties is negligible.

Reference to Fig. 2 indicates that a simple tracer con-

-
centration--mass balance yields for subchannel {
h6

dm; C. dx + wy (c. -Ci) dy = 0 (3


x

whe a

i=
We =
Woe
a4
(3)

Since diversion cross-flow is absent, Equation (31) becomes

VL
1G,
col a
wg (C.-C) -0 r3y

Similarly, we can write for subchannel

73° J
= + wy (c; - Ci) = 0

where, as assumed

"Mm "
mm. (34 J

Given the following initial conditions. for x = 0

7
A
Cc, C,- 0 (3° )

the solution of Equations (32) and (33) is:

50 dn (1- 26Si) (3¢ )


47
Fig. 9 illustrates the situation.

Fioure 9

wm; | C | \
—_—r= Ll
> Wu,

5,
i—mn=0

|
Equation (36) may be further simplified. If the outlet
/

concentration of the accepting subchannel, oF , is small compared

C. , then

LN me C; a
w .. =
v7)
~

td / C.

This result provides a simple analytical tool for rapid data evalu-

ation in the laboratory.

All values of flow were calculated from the flow meter

calibgation curves (see Appendix A). In recording the Reynolds

number for a given run, the port flow meter value was arbitrarily

selected as the "correct" reading. This was done to be consistent

with the experimental procedure (i.e., the flow was adjusted on


43
the port meter to a selected value, perhaps 50%, the starboard

meter did not necessarily read exactly the same value when the system

was balanced, in terms of the meter scale, but may have differed

slightly, say 51%), although the two flow rates were adjusted to be

equal. Appendix D contains the Reynolds number conversion factor

used in the data analysis.

B. Resul rs

The collected data of this experimental investigation

are presented in Appendix E in tabulated form. The data were

evaluated by the use of Equation (37), and are presented in a non-

dimensionalized form for comparison to other work in the field.

Fig. 10 is a plot of the collected data.

The line in Fig. 10 is drawn to indicate the probable

data trend; it is not a definitive result, and for this reason an

equation of the line is not presented. Also, the results have not

been "corrected" for entrance effects, since comparable experimental

runs were not made with a shorter window. Walton's [10] results,

as ''corrected" in Appendix B, are also plotted in Fig. 10 for

comparison.

The significant divergence of the lines in Fig. 10

indicates two factors in this author's data evaluation. The first

is the previously mentioned lack of "correction" for entrance

effects. The second factor is that the data used by this author

were evaluated by a simple two subchannel model. This yields a

more conservative result for the measured turbulent interchange.


490

Figure 10

Experimental Results

!
J)
Pat
4
=44

Walton

/
7

N
ta Fal

A
) \ /, A

v
2 A
/
A ’y

A
A
y
Lic

A
A
\

i 1 »

| {1“4 105
Qc
50
The actual mixing process involved six subchannels; there-

fore, the subchannel concentration of interest was diluted by two

peripheral subchannels. Galbraith [15] compensated for this factor

by solving, simultaneously, a set of eight mixing equations for his

multi-channel test section. This procedure corrected, to some

degree, the errors caused by the experimental measuring technique.

Although the data trend indicated in Fig. 10 seems to

compare favorably with Walton's results, there were many cases in

which the data deviated significantly from expected results.

Reference to Table E-I will indicate to the reader two important

inconsistencies in the data. The first is the significant divergence

of the mixing results for paired runs (i.e., for the same flow

conditions, but alternate tracer injection). This indicates that

diversion cross-flows were present during some of the experimental

runs. The author believes that this problem was due to the mano-

meters used for pressure balancing during testing.

As described earlier, in section IV-A, these manometers

are inclined draft gauges, for use with air. When 2.96 specific

gravity fluid is used in place of the design fluid (0.834 specific

gravity), the scale calibration in hundredths of an inch of water

no longer applies, but is reduced in accuracy by approximately one

fourth. The limit of accuracy was reduced to approximately four

hundredths of an inch of water. However, it was also found that

the reliability of the meniscus level upon fluctuations in readings

was much less than the corrected scale accuracy. On this basis,

his author estimates the accuracy of these inclined manometers


51
for this application to be plus, or minus, one tenth of an inch of

water. This is an unacceptable limit of accuracy and is the probable

cause of the divergence in some of the paired test runs.

The second source of inconsistencies in the data was

of a more serious nature. It became apparent that the response of

the instrumentation to variations in the injected salt concentration

was not behaving in a consistent manner. Although the instrument

employed could detect very small concentrations of sodium chloride,

it could not resolve small variations accurately. This deficiency

was compounded by extreme sensitivity of the measured conductivity

to temperature fluctuations and the natural fluctuations in the back-

ground value of conductivity. This inability to obtain reliable

data resulted in an abandonment of the experimental work envisioned

(i.e., shorter window length measurements for entrance effect

'correction' and traversing pitot tube probe measurements for con-

centration gradients across the subchannel gaps) in favor of in-

vestigating alternative instrumentation systems.


52

VI. CONCLUSION

A. Summary

This study has centered on providing a firm basis for con-

tinued experimental investigation into the mechanisms of turbulent

interchange. Test section construction delays and instrumentation

problems sharply curtailed the scope of this investigation. The

data obtained for mixing in a bare triangular array (P/D = 1.05)

seemed to compare favorably with other work in the field; however,

it was not considered to be sufficiently reliable to draw any

definitive correlation from.

B. Recommendations for Future Work

With specific reference to this investigation, two problem

areas are currently being dealt with. The manometers used for dif-

ferential pressure balance are being replaced with Dwyer ''magne-

helic" differential pressure gauges. For a full scale deflection

of five tenths of an inch of water, these gauges are accurate to

within two per cent (plus or minus one hundredth of an inch of

water). It is anticipated that these gauges will solve the pres-

sure balance problem encountered.

Various alternative instrumentation systems are being

explored for replacement of the conductivity technique. Spec-

trophotometry, in conjunction with a visual dye tracer, was investi-


-

33
3

gated and is currently being evaluated with respect to possible

pollution problems incurred by use of such dyes (see Appendix F).

Atomic absorption spectrophotometry and spectrofluorometry are two

possible techniques of extreme sensitivity and accuracy for use in

tracer studies of this kind. The spectrofluorometry technique

has the added advantage of continuous, on-line, capabilities.

The following areas of investigation are suggested for

‘uture work in the mixing field:

1. A systematic investigation of turbulent inter-

change in triangular array rod bundles remains to be done.

T'his research should focus on the variation of observed

mixing as a function of a range of pitch-diameter ratios

and Reynolds numbers.

2. Concurrent with the work mentioned above, a de-

tailed investigation of local concentration gradients ac-

ross subchannel gaps is necessary to distinguish develop-

ing mixing effects from fully developed values. Tra-

versing pitot tube probes will be incorporated in further

studies done in continuation of this project. This

technique provides a means of studying this problem in

detail and is a considerable improvement over the bulk

sampling techniques employed in the past.

3. Laminarization, secondary flows, and eddy dif-

fusivities are topics of special interest which will

require detailed investigation before the mixing process

can be completely characterized.


34

REFERENCES

l. Rogers, J.T., and Todreas, N.E., "Coolant Interchannel Mixing


in Reactor Fuel Rod Bundles: Single Phase Coolants," Heat Transfer
in Rod Bundles, ASME, New York, 1968.

2. Rowe, D.S., "COBRA-III: A Digital Computer Program for Steady


State and Transient Thermal-Hydraulic Analysis of Rod Bundle Nuclear
Fuel Elements," Report BNWL-B-82, 1971.

3. Rohsenow, W.M., and Choi, H.Y., Heat, Mass, and Momentum Transfer,
Chap. 8, Prentice-Hall, Inc., New Jersey, 1961.

4. Todreas, N.E., Ramm, H., and Johannsen, K., personal communi-


cation,tobepublished.
5. Coates, D.E., "Inter-channel Mixing and Coolant Temperature
Distribution in Seven and Nineteen Element Fuel Bundles," CGE
Report X0710001R, November 1960.

6. Moyer, C.B., "Coolant Mixing in Multirod Fuel Bundles," Riso


Report No. 125, July 1964 (issued 1966).

7. Elder, J.W., "The Dispersion of Marked Fluid in Turbulent Shear


Flow," Journal of Fluid Mechanics, 5, 1959, p. 544.

8. Todreas, N.E., and Wilson, L.W., "Coolant Mixing in Sodium


Cooled Fast Reactor Fuel Bundles,' WASH-1096, April 1968.

9. Petrunik, K.J., "Turbulent Mixing Measurements for Single Phase


Air, Single Phase Water, and Two Phase Air-Water Flows in Adjacent
Rectangular Subchannels,'" M.S. Thesis, University of Windsor, 1968.

10. Walton, F.B., "Turbulent Mixing Measurements for Single Phase


Air, Single Phase Water, and Two Phase Air-Water Flows in Adjacent
Iriangular Subchannels," M.S. Thesis, University of Windsor, 1969.

11. St. Pierre, C.C., "SASS Code I, Subchannel Analysis for the Steady-
State,' APPE-41, 1966.

12. Rowe, D.S., and Angle, C.W., ''Crossflow Mixing Between Parallel
Flow Channels During Boiling, Part II: Measurement of Flow and
Enthalpy in Two Parallel Channels,’ BNWL-371 PT2, 1967.
55
13. Rowe, D.S., and Angle, C.W., "Crossflow Mixing Between Parallel
Flow Channels During Boiling, Part III: Effect of Spacers on Mixing
Between Two Channels," BNWL-371 PT3, 1969.

14. Rowe, D.S., "A Mechanism for Turbulent Mixing Between Rod Bundle
Subchannels,'" ANS Transactions, 12, 1969, p. 805.

15. Galbraith, K.P., "Single Phase Turbulent Mixing Between Adja-


cent Channels in Rod Bundles," Ph.D. Thesis, Oregon State University,
1971.

16. Galbraith, K.P., and Knudsen, J.G., "Turbulent Mixing Between


Adjacent Channels for Single Phase Flow in a Simulated Rod Bundle,"
AIChE Reprint 19, 1971.

17. Singh, K., and St. Pierre, C.C., "Single Phase Turbulent Mixing
in Simulated Rod Bundle Geometries," Canadian Journal of Mechanical
Engineers, Vol. 1:2, June 1972.

18. Rogers, J.T., and Rosehart, R.G., "Mixing by Turbulent Inter-


change in Fuel Bundles: Correlations and Inferences,'" ASME, 1972.

19. Rogers, J.T., and Tarasuk, W.R., "Inter-subchannel Coolant


Mixing in Close-packed Reactor Fuel Bundles, Part I: Natural Mixing,"
CGE Report R64CAP29-1, 1966.

20. Rogers, J.T., and Tarasuk, W.R., "A Generalized Correlation for
Natural Turbulent Mixing of Coolants in Fuel Bundles," ANS Trans-
actions, 11, 1968, p. 346.

21. Rosehart, R.G., and Rogers, J.T., "Turbulent Interchange Mixing


Between Subchannels in Close-packed Nuclear Reactor Fuel Bundles,
Part I: Single Phase Coolant Generalized Correlation,' CGE Report
R69PP1, 1969.

22. Ingesson, L., and Hedburg, S., "Heat Transfer Between Subchan-
nels in a Rod Bundle," Fourth International Heat Transfer Conference,
Paper FC7, 11. 1970.

23. Van Der Ros, T., and Bogaardt, M., "Mass and Heat Exchange
Between Adjacent Channels in Liquid Cooled Bundles," Nuclear Engineer-
ing and Design, 12, 1970, pp. 259-268.

24. Johannsen, K., "Heat Transfer Analysis of Reactor Fuel Rod


Bundles," M.I.T. Nuclear Engineering Department Notes, 1973.

25. Dwyer, 0.E., "Recent Developments in Liquid Metal Heat Transfer,"


Atomic Energy Review, IV, No. 1, IAEA, Vienna, 1966.

26. Ramm, H., and Johannsen, K., "Radial and Tangential Turbulent
Diffusivities of Heat and Momentum Transfer in Liquid Metals,"
Progress in Heat and Mass Transfer, Vol. 6, 1973, pp. 45-58.
56
27. Nikuradse, J., "Turbulente Str8mung in nicht kreisf8rmigen
Rohren,'" Ing.-Arch. 1, 1930, pp. 306-332.

28. Singh, K., "Air-Water Turbulent Mixing in Simulated Rod Bundle


Geometries," Ph.D. Thesis, University of Windsor, 1972.

29. Sangster, W.A., "Calculation of Rod Bundle Pressure Loss."


ASME 68-WA/HT-35, December 1968.
57

APPENDIX A - Flow Meter Calibration

The flow meters were calibrated using standard laboratory

procedures. The two main channel meters are Fischer & Porter Model

10A3567 units rated at 9.5 GPM. Interchangeable floats were used

to obtain more accurate regulation at low flows. These floats

(F & P GSVGT-64) derated the units to 6.0 GPM maximum flow. The

collected data were evaluated by the least squares method to obtain

equations of the lines. These results are presented on the ac-

companying figures and were used in all data analyses. Reproduci-

bility of readings was generally within one per-cent.

The calibration curve for the tracer injection flow meter

was supplied by Fischer & Porter with the unit. This unit is the

F & P Model 10A1017-A rotameter. Sample readings confirmed the ac-

curacy of the calibrating curve to within plus or minus two per-cent.


58
Figure A-1

Port Flow Meter Calibration

3PM m=1b/min

EY)
~~
i

-—
a)

- 50

“50

C40)

F & P Model 10A3567


J)
Rated Flow 9.5 GPM

Equation of Line
Wa
} m= 0.79 x + 1.53

"

10
["

20 40 60 80
in siwiidicirirommiil—————ese
x = %Z rated flow
50

Figure A-2

Stbd. Flow Meter Calibration

5PM
wo
m=1b/min

v9

at )

30

91)

40

F & P Model 10A3567


< 3 Rated Flow 9.5 GPM

Equation of Line

m = 0.82x - 1.02

4)
y -

Lo,
20 40 60 80
r
.
wbiiemiviisomminsimiriomonsimemi——. ~ |
Xx = % rated flow
60
Figure A-3
Port Flow Meter Calibration
T T TT T
GPM m = 1b/min

1
oJ

’ ~
«1 J

LY

20

v=

F & P Model 10A3567


Rated Flow 6.0 GPM

0)
Equation of Line

m= 0.486x + 0.042

v 20 40 60 80
—_— ef
x = 7% rated flow
61
Figure A-4
Stbd. Flow Meter Calibration
GPM Im = 1b/minTt TTTTTT

af 1

*g5

F & P Model 10A3567


Rated Flow 6.0 GPM
IC
Equation of Line

m = 0.486x + 0.316

| §

JN

J
20 40 60 80
JIS I[NTEESEE
= 9% rated flow
Scale reading
tT TT

yd

- 20

i)

r—

de by
oO =
s3 09
cs
HOR
= o
G >
i i

2 (UF

10
F & P Model 10A1017-A
0
Tri-Flat Rotameter 0)
Ft
bdo

20 40 60 80 100 120
nl————————a——————— J. 1 | 140 160
3
ce/min |
63

APPENDIX B - Analysis of Walton's Data

Walton [10] used a simple mass balance in analyzing his

data. Based on the assumptions: (1) tracer concentrations do not

affect physical properties significantly, (2) tracer interchange does

not significantly reduce donor channel concentration, and (3)

diversion cross-flow is absent, then the turbulent interchange may

be expressed approximately as

wij - m Cj (E-1)
LC.
A more accurate approach, by a differential method, yields

4 vongphft-2] ’
C,
(FE-2)

When C; is very small compared to C. , the difference in

results using the two approaches is less than the experimental error

incurred.

Tables B-I and B-II are reevaluations of Walton's data

for single phase water and single phase air, respectively. The runs

are paired according to similarity of Reynolds number and injection

subchannel, but with different window lengths. The reader can


4

observe the nondimensionalized mixing, wf y , for each window

length and the "corrected" values (last column). The entrance ef-

fects were subtracted by converting the given data into "absolute"


4

mixing values (lb/hr versus 1b/hrft), taking the difference, and

dividing by the differential mixing length (0.75 ft).

In many cases, especially for Re less than 10000, the

reported mixing for the shorter window (09) exceeds that of the longer

(18). This can be attributed to the wake effect of the window

opening and the developing concentration gradient between subchannels,

which tend to raise the turbulent interchange level over that of

the natural mixing value (compare to flow scattering of spacer grids).

Fig. 6 is a plot of the "corrected" values. The results of this

analysis are not absolute in any sense, but indicate problem areas

in mixing data reduction and evaluation.


65

Sample Calculation

From Table III.2 of Walton's thesis [10], two runs

are selected, of corresponding Reynolds number and Channel injection,

but with differing window lengths. The mixing rate as recorded in

Ib/hr-ft is converted to 1b/hr by multiplying by the window length.

01R18 13.480 x 1.5 = 20.20 1b/hr

01RO09 17.276 x 0.75= 12.95 1b/hr

The shorter window length value is subtracted from the longer, giving

the approximate mixing in the last half of the longer window. This

is converted into a lb/hr-ft value by dividing by the differential

mixing length.

20.20 (01R18) - 12.95 (01R09) = 7.25 1b/hr

7.25 _
552 = 9.66 1b/hr-ft

Dividing by the viscosity ( 4 = 2.731 1lb/hr-ft at 60°F for water)

y ields the "corrected" dimensionless mixing.

9.66 _ Cw:
5 731 = 3:04 =

Note that this value is smaller than the reported values (Table B-I).
AG
Table B-1

Walton's Single Phase Water Data

Run # Re
Recorded Abs. Mix AM AM/1, Corrected
Ww’ /K 1b/hr 1b/hr 1b/hrft w’/ x

01L18 1983 5.32 21.8 11.0 14.68 5.36


01L09 1965 5.30 10.8

01R18 1983 4.94 20.20 7 25 9.66 3.54


01R09 1965 6.31 12.95

O03L18 2965 7.31 30.0 17.0 22.7 8.32


021.09 2905 6.34 13.0

O3R18 2965 5.44 22.3 6.1 8 15 2.28


02R09 2905 7.90 16.2

O5L18 3953 7.82 32.0 15.8 21.1 7.72


03L09 3874 7.88 16.2

O5R18 3953 6.44 26.4 11.15 14.85 5.44


03R09 3874 7.44 15.25

07L18 4956 7.45 30.5 18.05 24.1 8.82


041.09 4871 6.07 12.45

O7R18 4956 8.00 32.7 18.3 24 4 8.93


04R09 4871 7.02 14.4

09L18 5668 7.30 29.9 16.45 21.9 8.03


05L09 5555 6.66 13.45

O9R18 5668 8.24 33.70 21 90 29.2 10.7


J5R09 5555 5.76 11.80
67
Table B-II

Walton's Single Phase Air Data

Run # Re
Recorded Abs. Mix AM AM/L Corrected
w /y 1b/hr 1b/hr 1b/hrft w /y

01L18 5000 7.36 0.486 0.190 0.253 5.76


01L09 4766 8.98 0.296

O01R18 5000 9.26 0.611 0.297 0.396 a.00


01RO09 4766 9.50 0.314

O3L18 9862 14.95 0.988 0.498 0.664 15.10


031.09 10307 14.80 0.490

O3R18 9862 11.70 0.772 0.314 0.419 9.53


03R09 10307 13.90 0.458

06L18 19792 30.1 1.99 1.198 1.600 36.4


04L09 20652 24.0 0.792

06R18 19792 24.7 1.630 0.868 1.16 26.3


04R09 20652 23.1 0.762

09L18 49042 115 5.07 3 N5 4.06 92.2


071.09 53499 61.1 2 02

O9R18 49042 56.3 3.72 1 71 ) TR 51.8


07R09 53499 61.0 2.01

10L18 79629 111.0 }


10R1S8 79629 84.0 (No comparable data at higher Re for shorter
window)
68

APPENDIX C - Concentration Curve

To verify the linear variation of the conductivity readings

as a function of tracer concentration, a set of data was taken

throughout the concentration range of interest. This data is

tabulated in Table C-I. At each concentration, three measurements

were made. All samples were taken at 77°F to eliminate temperature

compensation corrections (the conductance bridge scales are based

on a 77°F reference temperature).

The multiple readings at each concentration served as a

check on the reproducibility of the measured conductivity. In

most cases, reproducibility of readings was within +1 per cent.

This data also provide an estimate of the sensitivity of the measuring

system. Concentration changes on the order of one part per million

salt in water are detectable. Fig. C-1 clearly indicates that the

readings are indeed linear, within the range of experimental accuracy.


69

Table C-I

Concentration Curve Data

NaCl Reading #1 Reading #2


2 x10~4 mhos x10~4 mhos Readingmhos
x10”
#3

Background 2.24 2.26 2.27


Ni 2.29 2.29 2.30
0.2 2.31 2.32 2.32
0.3 2.32 2.32 2.32
0.4 2.35 2.34 2.34
0.5 2.37 2.37 2.38
0.6 2.38 2.38 2.38
0.8 2.42 2.43 2.43
1.0 2.46 2.46 2.47
L.5 2.56 2.55 2.55
2.0 2.66 2.66 2.65
3.0 2.80 2.82 2.82
4.0 3.02 3.00 3.00
5.0 3.19 3.19 3.20
6.0 3.40 3.34 3.36
8.0 3.68 3.72 3.68
10.0 3.06 4.02 4.08

All readings taken at 77°F

Based on salt dissolved in 200 lbs. of water


Figure C-1
Concentration Curve

1.0 40
4.5 31.0

X 10”% mhos g/100 lbs. water

4.0

Zin
gg
5

2.6 -

re >
I

La

L

A y

Y T
J

A _—
aA -
_
/ 5

A ~~

2 7

a,
7
4
1 "NF i " ~

—~
71

APPENDIX D —- Reynolds Number Conversion Factor

In order to calculate the Reynolds number in the test

section subchannels it was first necessary to determine the flow

split. Fig. D-1 illustrates the subchannels involved.

"our 2 D1

Diameter = 6.25 x 10>ft


\ /
\ /
Pitch = 6.56 x 1072 ft _,
2
\
. / Area #1 = #2 = 3.3 x 10 = ft
Du, = 1.29 x 1072 ft
Dy, = 1.24 x 1072 ft

Follwing the method outlined by Sangster [29],

A ~ Ae, = A Dy (D-1)

wh — Tea

— 2

A .c a
tw
tL ev Tr 2)
-

D, 2,
vor Ap, L , and constant, the flow is proportional as
La

2
nD © ee Dag A (L-3)

Ce

A
A suitable expression must be chosen for the friction factor,

The Blasius relation is used


79

£ _ 0.316 A” .
0.25 D, (T 4)
Re (=>)
since it can be shown that the final result is relatively insensitive

to the value of the exponent of the Reynolds number chosen, within

the range of correlations suggested.

For constant values of fluid properties in the sub-

channels, Equations (D-3) and (D-4) combine to yield

mm
2-m _
a. D. {+m (D-5)

Since the total channel flow is the sum of the three subchannel

flows.

wm = my,+ cm, LEE


\ 3)

the flow in subchannel #1 is

mor (. 7)
22 D 1t+m
J———

1+ 2 a1 ( Du,ie ) 2
Substituting the appropriate values for the unknowns vields

rn 0.34 m w 3)

I'he Reynolds number then becomes


{5

Pe, = {97 x for (7-9)

This value was used in all data analysis.


74

APPENDIX E - Mixing Data

[he following coding is used for data collection and analysis,

149 presented in Table E-I:

Run # - odd numbers refer to port side tracer injection,


even numbers to the starboard side

Flow Rate - given in 1b/hr for one subchannel at the


window

Iracer Concentrations - Cg, refers to the initial donor


channel tracer concentration, C' to the exit con-
centration of the accepting channel; in each case
the common background value has been removed.

Mixing Rate - the mixing length of the window used was


I 27 f+.
75
Table E-I

Mixing Data

Flow |
Concen- Mixing ’
Run Rate Temp
#___ 1b/hr °C Re Tracer
Co
trations
_ © ~
Rate
1b/hrft
WwMif
——

001 273 24 4760 0.77 0.08 22.3 9.96


002 273 25 4890 0.91 0.08 18.9 8.67
003 193 3450 1.01 0.03 4.51 2.07
004 193 3450 1.09 0.04 5.56 2.55
J05 354 27 6680 0.03 11.96 5.78
006 354 29 6950 0.035 12.55 6.31
207 515 9240 0.04 17.3
708 515 9240 0.03 9.8
11800 0.035 12.05
11800 0.03 9.65
15000 0.01 3.73
15000 0.01 3.78
013 996 17400 0.04 20.4
014 996 17400 0.05 24.3
015 1158 23 19700 0.045 16.0
016 1158 23 19700 0.06 21.7
017 515 25 9240 0.06 11.3
018 515 25 9240 0.03 5.37
J19 515 26 9500 0.02 2.14
020 515 26 9500 0.03 3.14
021 515 9500 1.75 0.055 6.04
022 515 9500 1.85 0.05 5.24
023 675 12100 1.82 0.02 2.69
024 675 12100 1.79 0.06 8.17
025 675 12100 1.96 0.08 21.8 10.0
026 675 12100 1.80 0.06 17.8 8.16
027 837 15000 1.34 0.07 34.5 15.8
028 837 15000 1.41 0.07 32.7 15.0
029 1000 17900 1.74 0.05 22.6 10.37
030 1000 17900 1.69 0.04 18.6 8.52
76

APPENDIX F - Dye Injection Investigation

The use of a visual dye in connection with a spectro-

photometric instrumentation system is currently being investigated

in terms of possible environmental effects. A means of estimating

the dye concentration necessary for satisfactory operation of the

spectrophotometer is provided by use of Fig. F-1 [4]. The figure

indicates how the donor subchannel average concentration (or temper-

ature) is depleted as a function of the entrance length available

for mixing. For the geometry of this study and a Reynolds number

of approximately 20000, to detect in the parts per million range in

subchannel 4 , the concentration of subchannel ¢ (the donor) must

be approximately 100 ppm. A final assessment of the use of a visual

dve as a tracer awaits determination of possible adverse environmen-

tal effects upon discharge to the drainage system.


Figure F-1

Normalized Donor Subchannel Temperature Development [4]


1.000"
P/D = 1.05

1990J\ Pr
Re
=1
= 28,800

"ry-
Yo

— .970

T
960

950 ~

940

230) ~

100 {1
200
—7, -

You might also like