You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/222306774

Correlation of Sauter mean diameter and critical heat flux for spray
cooling of small surfaces

Article  in  International Journal of Heat and Mass Transfer · November 1995


DOI: 10.1016/0017-9310(95)00046-C

CITATIONS READS

213 417

2 authors, including:

Issam Mudawar
Purdue University
267 PUBLICATIONS   15,368 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Progress of Boiling Two-Phase Flow and Heat Transfer View project

Flow boiling in micro-channel heat sink View project

All content following this page was uploaded by Issam Mudawar on 28 January 2018.

The user has requested enhancement of the downloaded file.


Int. ,L Heat Mass Transfer. Vol. 38, No. 16, pp. 2985-2996, 1995
~ ) Pergamon Copyright © 1995 Elsevier Science Lid
Printed in Great Britain. All rights reserved
0017-9310/95 $9.50+0.00

0017-9310(95)00046-1

Correlation of Sauter mean diameter and critical


heat flux for spray cooling of small surfaces
K U R T A. ESTES and ISSAM M U D A W A R t
Boiling and Two-Phase Flow Laboratory, School of Mechanical Engineering, Purdue University,
West Lafayette, IN 47907, U.S.A.

(Received15 July 1994 and infinalform 12 January 1995)


Abstraet--F',xperiments were performed to understand better nucleate boiling heat transfer and critical
heat flux (CHF) for full cone sprays. The effects of spray nozzle, volumetric flux, subcooling and working
fluid were investigated. Dense sprays greatly reduced evaporation efficiency, and their boiling curves
exhibited an unusually small increasein slope upon transition between the singlephase and nucleate boiling
regimes. Sa~ater mean diameter (SMD) data were successfullycorrelated for fluids with vastly different
values of surface tension. This correlation was based upon orifice diameter and the Weber and Reynolds
numbers of the orifice flow prior to liquid breakup. Also developed was a new CHF correlation which
accurately predicted data for FC-72, FC-87 and water. This correlation shows a strong dependence of
CHF on volumetric flux and Sauter mean diameter. It is shown that by combining the correlations for
CHF and SMD it is possible to predict accurately CHF for full cone sprays without having to conduct
expensive and laborious drop sizing measurements for each individual nozzle.

1. INTRODUCTION While one would intuitively attempt to correlate


CHF data with respect to drop velocity, it is important
Despite the importance of sprays in many low-tem- to recognize that spray cooling is a commutative effect
perature, high heat flux situations, little information of many drops impacting the heated surface. Liquid
exists in the literature on the heat transfer mechanisms arrives at the surface with a volumetric flux Q" which
associated with spray cooling. Investigators have equals the volume flow rate striking an infinitesimal
suggested differenl parametric trends associated with portion of the impact area divided by the area of the
both nucleate boiJLing and CHF (critical heat flux), same portion. Different sprays can possess equal mean
but these trends were sometimes contradictory to one drop velocities but with vastly different volumetric
another and, to the most part, limited to a particular fluxes. Failure of drop velocity to account for the
nozzle and a single working fluid. commutative effects of multiple drop impact is why
Sprays are generally more difficult to characterize drop velocity is not recommended for correlating
than other boiling systems. Heat transfer correlations spray CHF data. In fact, all attempts at correlating the
for jets, for example, can be easily developed using present CHF data with respect to mean drop velocity
well-defined characteristic lengths (heater size, jet instead of volumetric flux proved very unsuccessful.
diameter) and characteristic velocity (jet velocity). The present study complements the study by Mud-
Choice of scaling parameters for sprays, on the other awar and Valentine [2] which provided dimensionless
hand, is complicated by the absence of a coherent fluid spray heat transfer correlations based on water data
flow downstream from the spray nozzle. As liquid obtained over a broad range of flow rates and for
breakup takes place, drops are formed which acquire drastically different spray types, including full cone,
different diameter:~, velocities, and trajectories, ren- hollow cone and flat sprays. CHF data were correlated
dering the use of' global scaling parameters ques- with respect to the local volumetric flux, Q", and
tionable. In fact, much of the difficulty in using exist- Sauter mean diameter (SMD), d32.
ing heat transfer data for sprays stems from the
contradictory recommendations by different inves- q'~ (pf,prQ"2d32,pfCp.fATsubI
tigators concernin;g which characteristic velocity and pgQ"hfg f \pg tr p~gfg /. (1)
characteristic length one should use in correlations.
For example, while most studies suggest spray CHF In a comprehensive review of the spray boiling
data be correlated with respect to the spray's volu- literature [6], it was determined that most studies,
metric flux and drop size [1-4], others recommend including Mudawar and Valentine's, were conducted
reducing the data relative to mean drop velocity and with only one working fluid, precluding any definitive
heater size [5]. assessment of the impact of key spray parameters,
such as surface tension, on the reported trends. The
tAuthor to whom correspondence should be addressed. present study attempts to: (1) develop an under-

2985
2986 K.A. ESTES and I. MUDAWAR

NOMENCLATURE

Cp specific heat at constant pressure AT~ temperature difference between heater


do nozzle orifice diameter surface and spray liquid, Tw-- Tf
d~2 Sauter mean diameter (SMD) AT~.b liquid subcooling, T~at - Tf
H nozzle-to-surface distance Tw heater surface temperature
hfg latent heat of vaporization Tf spray inlet temperature _ _
L heater length We spray Weber number, pfQ"2d32/a
P pressure We uo Weber number based on orifice flow
AP pressure drop across spray nozzle conditions.
q,, heater power divided by heater surface
area
Greek symbols
q~ heater power at CHF divided by heater
q evaporation efficiency
surface area
0 spray cone angle
q"p local critical heat flux defined at a point
p viscosity
q~ryout heat flux associated with complete
p density
evaporation of liquid (dryout)
qtt, cr surface tension.
dimensionless CHF
O total volumetric flow rate of spray
Q" local volumetric flux Subscripts
Q" volumetric flux averaged over spray a ambient (air or vapor)
impact area, Q/(TrL2/4) f liquid
Re ao Reynolds number based on orifice flow g vapor
parameters sat saturation
T temperature sub subcooling.

standing of nucleate boiling and CHF in sprays ; (2) upstream of the spray nozzle. The fluid was de-aerated
construct a universal CHF correlation for sprays using in the same loop prior to each test.
newly acquired data for Fluorinerts FC-72 and FC- Situated directly below the test chamber was the
87 along with the water CHF data obtained earlier by loop reservoir as shown in Fig. 1. A magnetically
Mudawar and Valentine. While the two Fluorinerts coupled centrifugal pump circulated the Fluorinert
share similar thermophysical properties (except for through the loop which consisted of insulated stainless
saturation temperature, 56 °C for FC-72 and 30 °C for steel and inert plastic components. Only a small frac-
FC-87), their surface tension is only about 17% of the tion of the total flow delivered by the pump (about
surface tension for water. The bulk of the new data 2.52 x 10-3 m 3 s -1 (40 gpm)) was actually routed to
was for FC-72 since the low saturation temperature the nozzle, the balance was bypassed to the reservoir.
of FC-87 resulted in pump cavitation and magnetic Two fiat plate heat exchangers were employed to
decoupling at high flow rates. remove the heat added to the liquid by the pump. The
Recently, Estes [6] determined that volumetric flux spray liquid was routed through a carbon filter before
is a maximum directly beneath the spray nozzle and being reheated by an in-line electrical heater, and
decays towards the outer edge of the spray impact either heated or cooled, as needed, via a third fiat
area. A theoretical model was developed which accu- plate heat exchanger. Water supplied to this third heat
rately predicted the volumetric flux distribution across exchanger was circulated through a constant tem-
the heated surface. He also demonstrated that CHF perature bath; this enabled fine tuning of the spray
for square heaters can be maximized by adjusting the temperature. One of two rotameters with overlapping
nozzle-to-surface distance such that the spray impact flow-rate ranges was used to measure the flow rate
area just inscribes the square surface of the heater. prior to entering the nozzle, and the nozzle inlet pres-
Therefore, all the Fluorinert data presented in the sure was measured with the aid of a pressure gauge
present study were taken with the spray conforming situated just upstream of the nozzle. After exiting the
to this optimum configuration. nozzle and impinging upon the heater surface, the
evaporated portion of the flow was condensed within
the test chamber and recovered along with the unevap-
2. EXPERIMENTAL FACILITY
orated liquid through a large diameter drain at the
A two-phase flow loop was assembled which both bottom of the test chamber, leading directly to the
maintained the desired pressure within the test cham- reservoir.
ber housing the heater and facilitated precise con- As shown in Fig. 2(a), the test chamber housed the
ditioning of the fluid flow rate and subcooling heater module, spray nozzle assembly and a water-
Sauter mean diameter and critical heat flux for spray cooling 2987

Reflux
Condenser ~ _ a ~

# Gauge

I I
Test I I Flowmeters

Condensing Pressure
co, ~
~-_@ank Control

Rater ~ Immersion
Heat
Exchanger
~.._~ Condensing Heaters
F k,,,,,JI Coil
Re.c;ervoir Constant
In-line Temperature
Electrical Bath
Heater
}]eaters

Heat Heat I I Carbon


Exchanger Exchanger L ) ,Filter
Pump
Fig. 1. Schematic of flow loop.

cooled condenser The heater was machined from oxy- Table 1 gives the important parameters for the three
gen-free copper and was powered by a thick-film elec- nozzles.
trical resistor silver soldered to its underside. As
shown in Fig. 2(b), the heater rested upon a bracket
Spray boiling curves
which provided thermal insulation on all surfaces of
Boiling curves for FC-72 at 13 °C subcooling are
the heater excepl; the impact surface. Surface tem-
presented in Figs. 3(a)-(c) for nozzles 1, 2 and 3
perature was extrapolated from a thermocouple
respectively. At the low volumetric flow rates associ-
embedded beneath the surface, assuming one-dimen-
ated with nozzle 1, a sharp increase in slope seems to
sional heat conduction. The surface of the heater was
mark the commencement of nucleate boiling.
polished prior to each test.
However, at the larger flow rates associated with noz-
Located along a normal axis centered with respect
zles 2 and 3, this increase in slope is much less discern-
to the heater surface was a vertical tube to the end of
ible. In fact, CHF for both nozzles seems to occur
which different spray nozzles could be mounted. The
following only a short nucleate boiling heat flux span.
nozzle was raisecL or lowered to the desired precise
Figures 4(a)-(c) show boiling curves for nozzle 1 at
placement relative to the heater surface with the aid
three increasing flow rates that illustrate the effect of
of a digital micrometer translation stage bolted atop
subcooling. At the lowest flow rate, the nucleate boil-
the test chamber.
ing regime is quite evident at all three subcoolings.
Uncertainties in the measurement of pressure, flow
However, the nucleate boiling heat flux span gets
rate and heater power were estimated at 0.5%, 1.6%
smaller as subcooling and flow rate increase, becom-
and 1% respectively, and thermocouple measure-
ing virtually non-existent for the highest volumetric
ments were made with a _+0.2 °C accuracy. Heat loss
flow rate and highest subcooling. Evidently, the span
to the heater surroundings was determined numeri-
of the nucleate boiling regime is dependent on both
cally to be less than 4.5% and 1% during the single-
subcooling and flow rate, lower subcoolings and flow
phase and boiling regimes respectively.
rates being conducive to a broader nucleate boiling
heat flux range.
Figure 5 shows spray boiling curves for water
3. RESULTS AND DISCUSSION
adapted from ref. [2]. These curves clearly depict an
Three Spraying; Systems full cone spray nozzles, increase in both the single-phase heat transfer
designated as nozzles 1, 2 and 3, in order of increasing coefficient and CHF with increasing Q" or decreasing
flow capacity, were used to acquire the FC-72 data. d32. Also evident in the same figure is the convergence
2988 K . A . ESTES and I. M U D A W A R

Micrometer

ninr
(a) - - Translation
Stage

Spmy
Nozzle

7~ He=er

- - Cover
(Lexan)

S Condensing
Coil

I ora,n twato
(b) ChromeI-Alumel
(Type-K) Heater
Thermocouple (Oxygen-Free Copper)
16.51
in 0.81 mm
Diam. Hole
\ 12.70
Thick Film Resistor
Insulation
6,10 0.51 (G-7 FiberglassPlastic)

3.05, I - - - - - ~\\\',~ "///IVJ


////I// ®/ Insulation
(G-10 Fiberglass Plastic)

N
All dimensions in millimeters
Fig. 2. (a) Schematic of test section and (b) construction of heater module.

Table 1. Characteristics of the spray nozzles using FC-72

Spray mean Sauter mean


Orifice diameter Spray angle volumetric flux diameter (SMD)
do x t06 0 Q'-~× 103 d32 x 106
Nozzle (m) (degrees) (m 3 s - i m - z) (m)

762 55.8 16.6-52.4 110-196


1190 46.4 55.6-145 181-225
1700 48.5 86.3-216 182-214
Sauter mean diameter and critical heat flux for spray cooling 2989

(a)
. . . . . . . . i . . . . . . . . ! I
FC:72 . . . . . . . = ........ '
(a) FC-72
P = 1.03 bar (15 psia) Z; - 1 P = 1.03 bar (15 psla)
T~t = 57.3 °C
Tsat '= 57.3 °C ~. 100
100 i ,,,m=,,=13, c ,~ E:~ Q = 3.53x10Sm3/s ~
I (0•056 gpm) #]~r
Ic"F---"1 ,p~
CHF "-'~P" I 7

=1=" 10 Z~ ~ I Nozzle 1 =o" 10 ~-~-~ .1- - - I


~0" q;n
w~n~
0 13 85
io i,o:o.,
5.55
i-
[] 23 e9
I ~ [<o.,)I ,~
1 . . . . . . . . i
I~1 ,.,
(0.120) I
. . . . . . . . i
137
1 i i . . . . .i . .
ZI
i
33
i i • ..
93
I
10 100 10 100

AT w (°C) ATw (°C)


(b) ........ ' ........ ' (b) FC:72 . . . . . . . i . . . . . . . . ,
FC-72 ]
P = 1.03 bar (15 psia)
P = 1.03 bar (15 psia) I
Ts=t,= 57.3 °C I T~t = 57.3 °C _ 41~_A,~;.
,oo ~u,=,~oc l ~ 100 Q = 5.55x10-Sm3/s
I (0.088 gpm) j ~
~"
~
~. ICHF---"I / ~.~ I~ ~-------/~-J

lo ~ ~ ] . =~ lo ~ - ~'~-~T;~:--I
I I~o.lq~n.I]
I I <o,'n)I W~/] •c w/crr~ I

I o I,o% I ,,/1 O 13 100 I

[] 23 123 ]
17.3
A 33 115 I
1 1 , , ...... ' ' ' ' ''"
10 100 10 100

AT w (°C) AT. (°C)


i . . . . . . . . i I
(c) (C) FC:72 . . . . . . . ' ........ '
FC-72 ] t P - 1.03 baor (15 psia)
P = 1.03 bar (15 psia) I Tsar = 57.3 C
• at -, 57.3 °C ,~
loo i A.r,=,= 13 oc I ;! 100
I
O = 7.57x10.Sm3/s
(o.120 gpm) ~ r

0
'o" 10 ~ I Nozzle3 =,:,- lo ofi~
I I m~o'l ~m ATsub
•C
q;"
W/cm~
]
I i O 13 137

I [] I ~:o~I - Z~
[]

33
23 127

1~
1 , I ~ l~:.~o,l :,, 1 , , ...... ' ........
10 100 10 100

AT w (°C) &T. (°C)


Fig. 3. FC-72 boil:Lng curves for three flow rates at 13 °C Fig. 4. FC-72 boiling curves for nozzle 1 for three subcoolings
subcooling for (a) nozzle 1, (b) nozzle 2 and (c) nozzle 3. at (a) Q = 3.53× 10 -6 m ~ s -~, (b) Q = 5.55× 10 -6 m 3 s -1
and (c) Q = 7.57× 10 - 6 m 3 s i.
2990 K.A. ESTES and I. MUDAWAR

1000

g-
E
¢,,)
100

10
10 100 1000
ATw(°C)
Fig. 5. Boiling curves measured by Mudawar and Valentine [2] for full cone water sprays.

of all the nucleate boiling data onto a single line, smaller volumetric flux and, to a lesser extent, greater
regardless of volumetric flux or drop diameter, and surface tension. A small volumetric flux reduces liquid
the relatively large slope (about 5.6) of the nucleate buildup, exposing the surface to direct impingement
boiling regime. Comparing these water boiling curves by the drops. Evaporation of a large percentage of the
to the curves for FC-72 given in Figs. 3(a)-(c) and spray liquid is manifest both in a larger heat flux span
4(a)-(c) clearly points to the sensitivity of the slope in in the nucleate boiling regime and a large increase in
the nucleate boiling regime to working fluid, volu- slope upon transition from single-phase liquid cooling
metric flux and subcooling. to nucleate boiling. This is illustrated clearly in Figs.
3(a)-(c) where the low flow rate, low Weber number
Evaporation efficiency cases exhibit a large increase in slope and the boiling
A parameter characterizing the efficiency of liquid curves with higher volumetric flow rates (higher
evaporation was employed to explain the unique Weber numbers) do not.
shape of boiling curve described in the previous Using extensive high-speed photographic analysis
section. This efficiency was defined as the percentage of individual water drops impinging upon a heated
of the total heat that could be removed by the spray, surface, Bernardin [7] showed that even at surface
sensible and latent, that was actually removed at C H F temperatures close to CHF, boiling does not com-
mence in the impinging liquid until the drop flattens
qm out into a thin liquid film and loses most of its radial
~/ ---- q~ryou~ × 1 0 0 %
momentum. Thus, the early stages of drop impact
q~ during the nucleate boiling regime are dominated by
x 100%. (2) single-phase heat transfer. The high, instantaneous
prQ"hfg (1 +cp,rATs,~b/hrg)
single-phase heat transfer coefficient associated with
Figure 6 shows efficiency vs spray Weber number for the droplet impingement seems to suppress any
FC-72 and FC-87 at all volumetric flow rates and nucleation. Only as the convection coefficient greatly
subcoolings and all three nozzles. Also shown are diminishes due to loss of liquid momentum do the wall
water data for full cone sprays measured by Mudawar cavities acquire enough superheat to begin boiling.
and Valentine. Figure 6 shows that evaporation This important mechanism sheds much light upon
efficiency is inversely related to the Weber number, the evaporation efficiency trends revealed in Fig. 6. A
decreasing from a nearly 100% efficiency for spray possessing a small volumetric flux (i.e. low spray
We < 10 -5 to less than 10% for We > 0.1. Water tests Weber number) can be classified as a light spray. Con-
had much smaller Weber numbers than the FC-72 and versely, a high volumetric flux (i.e. high Weber num-
FC-87 tests, resulting in more of the liquid impinging ber) spray carries the classification of dense spray,
upon the heated surface evaporating. A relatively Figure 7 compares nucleate boiling with those two
small Weber number (water data) corresponds to a types of sprays. In a light spray, the frequency of drop
Sauter mean diameter and critical heat flux for spray cooling 2991

FC87 FC72 Water


&Tsub-4 15"0 ~Tub=13 3300 AT,~=20 77*C
0 Noz~l • 45° Full C ~ e

100 [] Noz~e 2 • 3(P Full Cone


• ~o •
• • Non~3 • 15" FUllCone

c-
O

~ 10

1
106 105 104 10-3 10-2 101 100 101
Pf ~ 2 d =
O

Fig. 6. Evaporation efficiency vs spray Weber number.

Approaching
spraydrops

Impinging Boiling following


drop ~ • / dropspreading

Light Spray: Low We, high evaporation efficiency

O0
0 0 O
Impinging
O0 O, Boiling following
ig

Dense Spray: High We, low evaporation efficiency


Fig. 7. Schematic representation of boiling in light and dense sprays.
2992 K.A. ESTES and I. MUDAWAR

500 i . . .
FC-72
. .
(0.04 to 0.5 gpm) and subcooling from 13 to 33 °C
ATaub Nozzle
(°C) 1 I 2 I 3 P = 1.03 bar, Tsar= 57.3 0( using the three spray nozzles with overlapping flow-
~' 13 0 [] A rate ranges.
Figure 8 shows C H F values for FC-72 for all three
nozzles for the high and low subcoolings; inter-
:E 100 mediate values have been left out for clarity. C H F was
s found to increase with increasing flow rate for each
50 ' ........ nozzle at all values of subcooling. Nozzle 1 had larger
10"6 10-5 10 -4 C H F values than the two larger nozzles at the same
O (m3/s)
flow rate, highlighting the important effect of drop
. . . . . . . . i , , ...... i
0.1 1 size on CHF. Smaller drops possess a greater surface
Q (gpm) area to volume ratio than do larger drops ; therefore,
the smaller drops can utilize their sensible and latent
Fig. 8. CHF vs total volumetric flow rate for nozzles 1, 2 and
3 using FC-72. heat more effectively than the larger drops can. In the
present study, nozzle 1 produced much smaller drops
for a given flow rate than did nozzles 2 or 3. Likewise,
impingement upon the heated surface is low, leaving drops of nozzle 2 were slightly smaller than those for
much of the surface covered with fairly stagnant liquid nozzle 3 at the same flow rate, and, as indicated in
within which vapor bubbles can easily nucleate. Evap- Fig. 8, CHF values were greater for nozzle 2 than for
oration efficiency in light sprays is, therefore, very nozzle 3. Figure 8 also shows C H F increased with
high. Dense sprays, on the other hand, are char- increasing subcooling for each nozzle because of the
acterized by a high frequency of droplet impingement increased amount of heat the liquid could absorb prior
which suppresses nucleation over a large fraction of to evaporation.
the heated surface, resulting in a poor evaporation
efficiency. This description of drop impact during
4. SAUTER M E A N DIAMETER CORRELATION
nucleate boiling also explains why volumetric flux is
of much greater significance to characterizing spray The sprays tested in the present study were char-
heat transfer than drop velocity; for, while drop vel- acterized by a Phase Doppler Particle Analyzer
ocity affects the local heat transfer from the heated (PDPA). This non-intrusive technique enabled deter-
surface momentarily, it is the volumetric flux that mination of the spray hydrodynamic parameters with-
determines the commutative effect of multiple drop out disrupting the spray itself. As shown in Fig. 9,
impingement. In fact, all attempts at correlating the light produced by a 3 W argon-ion laser was split into
present spray CHF data relative to drop velocity pro- two coherent beams using a beam splitter. Measure-
ved very unsuccessful. ments were made at the control volume formed by
intersection of the two beams. The interference fringe
Critical heat flux pattern produced by light scattered by individual
A C H F database was collected using FC-72 for flow drops as they passed through the control volume was
rates ranging from 2.52 × 10 -6 to 3.15 × 10 -5 m 3 s -I captured by optical detectors. The phase difference

Transmitter ~ v "... - -..-"


Lens m - ~ - ........ -
. ~ " Measurement Volume

I
~ ~ (defined by the Intersection of
' ~ two laser beams)

Argon.lon laser Beam "" ~ ' - . .


(3 W) Spiltter S=.tered Ugi~ "
(detectable by the / / ~

Receiving ~ ~'-~---~

I Signal Pr_oees__sorI
and I
Microcomputer I
Fig. 9. Phase Doppler Particle Analyzer (PDPA) optics (adapted from [8]).
Sauter mean diameter and critical heat flux for spray cooling 2993

(a) 2500 (b) 2500


."[1 I Flud Water
2000

1500

~'1000 ~1000
a a
5OO 500 |IEll I ,o..le, I

2000 ..e- 2000


c -n
1500 1500

~'1000 1000
D
5O0 i Nozzle 2 500 I ,ezzle2 I

tk 8
2000 "E 2000

1500 1500

~'1000 1O00
a
50O Nozzle3 500
i ~ i ~ I d~196~n o
"1O0 200 300 400 500 1O0 200 300 400 500
Diameter OJm) Diameter/dam)
Fig. 10. Drop size histograms for nozzles 1, 2 and 3 using (a) FC-72 and (b) water.

between the signals from the different detectors was Therefore, a new correlation is developed using SMD
proportional to the size of the drop passing through data for the three full cone spray nozzles and two
the control volume. The detector signals were pro- fluids : FC-72 and water.
cessed and condhioned for data acquisition and analy- Lefebvre [9] proposed a systematic model to
sis using a microcomputer. describe the breakup of a sheet of liquid into spray
The spray's SMD was determined along the axis droplets that was successful at correlating SMD data
of the spray for' each nozzle at room temperature, for hollow cone sprays. He suggested the breakup was
Tf = 23 °C. Measurements made away from the axis caused by both aerodynamic forces downstream of
proved SMD w~.s fairly uniform. The measurements the orifice as well as turbulent fluctuations in the liquid
along the spray axis were taken at specific nozzle-to- upstream of the orifice. He recommended a general
surface distances and pressure drops. Drop sizes in the correlation for dimensionless SMD, ratio of SMD to
FC-72 sprays were measured at a distance of 45 mm thickness of the liquid sheet, as a function of a Weber
from the nozzle orifice which both ensured a fully number and a Reynolds number, both based on liquid
developed spray and precluded the multiple beam velocity and thickness of the liquid sheet.
scattering sometimes encountered in the measurement Since full cone spray nozzles do not produce a liquid
of very dense sprays. sheet prior to exiting the orifice, the characteristic
Figures 10(a) and (b) show, for each of the three length and velocity chosen for correlating SMD in the
nozzles, drop size histograms for FC-72 and water present study were the orifice diameter, do, and liquid
respectively, for the same pressure drop of 276 kPa velocity at the orifice, defined as (2AP/pf) °5. The
(40 psi). Generally, water produced slightly larger Weber and Reynolds numbers based on the orifice
drops than FC..72. Drop sizing was repeated for conditions were defined, respectively, as
different pressure drops, and curve fits were developed
to determine SMD for the pressure drop at which pa(2AP/pf)do
Wedo -- (3)
CHF was later measured. o"

A correlation for SMD would enable a designer to


and
predict heat tra~asfer performance of a given spray
without having 1:o perform the costly and time con- pf( 2 A P / pr) '/2do
suming optical drop sizing. Many correlations have Redo -- (4)
been developed which predict SMD for various spray
types, however, none is available for full cone sprays. where Pa is the density of ambient fluid (air or vapor).
2994 K.A. ESTES and I. MUDAWAR

FI
I. I
. Nozzle
IIdo- . 1 . Nozzle
. 0.762 . . 2 . Nozzle
. . 3. . Pa. . . Pt.
do= 1.19 do= 1.70 a X 10a /.z¢X 10e
]. [ ] (mm) (mm) (mm) (kg/m 3} (kg/m 3) (N/m) (kg/m.s)
t__~:~ ~ mE] ~ 1.20 168012.2 663
1.20 998 72.8 959
o
~ ~d32 / do= 3.67 (We°~sRedo) 0.259
~ ~ ' ~ ' ~ , ~ ~ ~. (meanabsolute error = 12.4 %)
"O
~ ~ _ ~ ~ ~ ~ +/-25%

0.1
10 4 10 s 10 6
We°~s x Red,
Fig. 11. SMD correlation.

The resulting correlation, Q", as the total spray volumetric flow rate divided by
the impact area. It is postulated that surface dryout
(t32 = 3.6 7[ Wela/o2Redo] -°259 (5) at CHF commences at the outer edge of the impact
do area, where volumetric flux is a minimum, since this
fits the SMD data for both FC-72 and water with a local dryout would reduce the fraction of the surface
mean absolute error of 12.4% as shown in Fig. 11. area available for cooling and increase the heat flux
Interestingly, the product of the square root of Weber within the impact area, enabling the dryout region
number and Reynolds number in equation (5) is ident- to propagate radially inward in an unstable manner.
ical in form to the product proposed by Lefebvre for Therefore, CHF at the edge of the impact area should
hollow cone sprays. govern CHF for the entire surface. Critical heat flux
at the edge was determined by assuming all of the heat
was transferred through the impact area,

5. CHF CORRELATION q"L 2 4


"
qm,p (~z/4)L2 -- 7zq". (6)
Mudawar and Valentine [2] correlated CHF data
for water using equation (1). Unlike the present Flu- Volumetric flux at the outer edge of the impact area
orinert data, their water data were measured by a was determined from the spray flux distribution model
circular test heater whose surface area was much smal- developed by Estes [6]
ler than the spray impact area. Essentially, both CHF
Q,,
and volumetric flux were measured at a single point
Q,~ - ~[1 + cos(0/2)] cos(0/2). (7)
in the water sprays ; hence, the respective designations
qnp and Q". Note that, for the water data, the mean
volumetric flux upon the heater's surface was equal to Figure 13 shows that the correlation
the local volumetric flux, Q" -- Q". This water spray
configuration is illustrated in Fig. 12(a), which also qm,p _ 2.3(Pf~°'3(PrQ"2d321-03'
includes a plot showing CHF for different full cone
pghfga" \pg/t \ ~ /
spray nozzles increasing with increasing volumetric × (1 +O'O019PfC°'fAr'ub~ognrg
/ (8)
flux and subcooling.
The CHF correlation form proposed by Mudawar
and Valentine for water data was employed to cor- fits the CHF data for water, FC-72 and FC-87 with a
relate the FC-72 and FC-87 CHF data obtained in mean absolute error of 12.6%. SMD for FC-87 was
the present study. Two modifications were needed to determined using the drop size correlation developed
allow the Fluorinert data to be correlated with the earlier in equation (5), as drop sizes for FC-87 were
water data using the same equation. As shown in Fig. not measured. Accurate prediction of the FC-87 data
12(b), the Fluorinert sprays were configured so that is proof that CHF for full cone sprays can be predicted
their impact area just inscribed the heater surface fol- using equations (5) and (8) without the need for
lowing the optimization guideline developed by Estes expensive drop-sizing equipment. Close inspection of
[6]. For the Fluorinerts, critical heat flux, q", was equations (5) and (8) also reveals which spray par-
measured as the heater power divided by the total ameters actually influence CHF. These parameters
heater surface area, and the mean volumetric flux, include thermophysical properties (Pr, Pg, O', hfg, Cpf),
Sauter mean diameter and critical heat flux for spray cooling 2995

(a)
distance to ensure complete breakup into drops. Poor
breakup could greatly reduce C H F relative to pre-
dictions based upon equation (8). Some of these effects
are evident in the lowest volumetric flux data in Fig.
13 for each of the nozzles used with FC-72 and FC-
87. The lowest volumetric flux for each nozzle cor-
responds to the lowest nozzle pressure drop required
for liquid breakup. These data show the largest devi-
ation from the correlation whereas the high volu-
metric flux data, which correspond to large pressure
drops and complete breakup, are more accurately pre-
dicted by the correlation.
1000 i
e &T=ub Water
:degrees: ('C) P=l.01bar, T~t=100°C
o 15 77
30 77
45 77
6. CONCLUSIONS
15 20-77
=E A Experiments were performed to understand better
o"
A nucleate boiling and C H F in sprays, and to develop
.... i
lO0 correlations for the spray's SMD and CHF, which
lO-4 10-3 10"2
Q,, ( m 3 s - l l m 2)
would enable accurate determination of C H F for
(b) different fluids and different full cone nozzles without
having to conduct expensive and laborious drop-siz-
ing measurements. The following are the key con-
clusions from the study.

(1) The shape of the boiling curve of sprays is mark-


edly different from that of other boiling systems.
Sprays with high volumetric flux (high Weber
number) show little increase in slope of the boiling
curve between the single-phase and nucleate boil-
ing regimes because of a suppression of nucleation
and reduced evaporation efficiency. Low volu-
500 i metric flux (low Weber number) sprays, on the
6T=ub Nozzle 1"(.;-7Z
P = 1.03 b a r , Ts.,~= 5 7 . 3 ° C
other hand, display a more pronounced increase
(*C) 11213
in the slope of the boiling curve because of a
Eo
13
33
OlOl
• •
A• higher evaporation efficiency.
(2) C H F increases with increasing flow rate and
:EIO0 : increasing subcooling; C H F is also greater for
=' S nozzles which produce smaller drops.
(3) SMD for full cone sprays is dependent upon ori-
50 " " " ' ' " ""
0.01 0.1 fice diameter and the Weber and Reynolds num-
Q,-"; (m3s-1 / m 2) bers based on the orifice flow conditions prior to
Fig. 12. (a) Mudawar and Valentine's CHF data for water liquid breakup, A dimensionless correlation was
[2] vs local volumetric flux and (b) present CHF data for developed which gives good predictions for fluids
FC-72 vs mean volumetric flux. with vastly different surface tensions.
(4) A correlation was developed which accurately
predicts C H F for water, FC-72 and FC-87 and
many different full cone nozzles over a wide range
of flow rates and subcoolings. This correlation
flow parameters (.~Tsub, AP, Q), orifice parameters
demonstrates that C H F is influenced by the ther-
(do, 0), and heater length (L). The important effect of
mophysical properties of the fluid (Pr, Pg, a, hfg,
nozzle-to-surface distance (H) is implicitly a function
cp.f), flow parameters (ATsub, AP, Q), orifice par-
of both heater size and spray cone angle for a spray
ameters (do, 0), and heater length (L).
which has been configured to optimize C H F ; one
whose impact area just inscribes the square heated
surface.
The reader is cautioned to avoid using the present Acknowledgments--Financialsupport for this work by IBM
is greatly appreciated. The authors also thank the 3M Com-
C H F correlation for sprays which are not hyd- pany for donating Fluorinert samples, and Spraying Systems
rodynamically fully developed. Sprays require a mini- Company for both donating spray nozzles and assisting with
mum pressure drop and minimum nozzle-to-surface the spray droplet sizing.
2996 K.A. ESTES and I. MUDAWAR

FC-~, p. t.r~bN, T~ .573 tC FC-87,p. 1,n~bw, T~ .30.7~C INMM,P. 1.01her.Tee=-100°(


No=~l Nozz~2 Nozzle3 Nozzle1 Sewralful conespraynozzles
o [] ~ e
Q" X103 16.6-52,4 55.6-145 86.3-216 16.6-49.7 0.600-0.218
m3s-1/m2
d32x 106 140°196 181.225 182-214 110-195 405-1350
m
&Tl~e 13-33 13-33 13-33 4-15
°C
1000 .....i i . . . . . . . i i .... i . . . . . .

q~,p

100 ." F l°' F, + o.oo,, ,, •


q: / / ~ ~ / " f ~ ~ +/- 30%

10
---L o J
~ . . . ~ . . . . . . . . . . . . . . :~ : : il'ror 26%,

1
10-1 10 0 101 10 2 10 3 10 4 10 s 10 s
p! Q"2d32

Fig, 13. CHF correlation.

REFERENCES 6. Estes, K. A., Critical heat flux in spray cooling and jet
1. S. Toda, A study in mist cooling, Trans. JSME38, 581- impingement cooling of small targets, Masters Thesis,
588 (1972). School of Mechanical Engineering, Purdue University,
2. I. Mudawar and W. S. Valentine, Determination of the West Lafayette, IN (1994).
local quench curve for spray-cooled metallic surfaces, J. 7. Bernardin, J. D., Intelligent heat treatment of aluminum
Heat Treating 7, 107-121 (1989). alloys: material, surface roughness, and droplet-surface
3. G. E. Totten, C. E. Bates and N. A. Clinton, Handbook interaction characteristics, Masters Thesis, School of
of Quenchants and Quenching Technology, pp. 239-289. Mechanical Engineering, Purdue University, West Lafay-
ASM International, Materials Park, OH (1993). ette, IN (1993).
4. I. Mudawar T. A. and Deiters, A universal approach to 8. D. D. Hall, A method of predicting and optimizing the
predicting temperature response of metallic parts to spray thermal history and resulting mechanical properties of
quenching, Int. J. Heat Mass Transfer 37, 347-362 (1994). aluminum alloy parts subjected to spray quenching, Mas-
5. C. S. K. Cho and K. Wu, Comparison of burnout charac- ters Thesis, School of Mechanical Engineering, Purdue
teristics in jet impingement cooling and spray cooling, University, West Lafayette, IN (1993).
Proc. National Heat Transfer Conf., pp. 561-567. Hous- 9. A. H. Lefebvre, Atomization and Sprays, Hemisphere
ton, TX (1988). Publishing Corporation, New York, NY (1989).

View publication stats

You might also like