You are on page 1of 15

This article was downloaded by: [University of California Santa Barbara]

On: 28 June 2013, At: 05:45


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number:
1072954 Registered office: Mortimer House, 37-41 Mortimer Street,
London W1T 3JH, UK

Experimental Heat Transfer:


A Journal of Thermal Energy
Generation, Transport,
Storage, and Conversion
Publication details, including instructions for
authors and subscription information:
http://www.tandfonline.com/loi/ueht20

A GENERAL HEAT TRANSFER


CORRELATION FOR
IMPACTING WATER SPRAYS
ON HIGH-TEMPERATURE
SURFACES
S. C. Yao & Timothy L. Cox
Published online: 10 Nov 2010.

To cite this article: S. C. Yao & Timothy L. Cox (2002): A GENERAL HEAT
TRANSFER CORRELATION FOR IMPACTING WATER SPRAYS ON HIGH-TEMPERATURE
SURFACES, Experimental Heat Transfer: A Journal of Thermal Energy Generation,
Transport, Storage, and Conversion, 15:4, 207-219

To link to this article: http://dx.doi.org/10.1080/08916150290082649

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/


terms-and-conditions

This article may be used for research, teaching, and private study
purposes. Any substantial or systematic reproduction, redistribution,
reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden.
The publisher does not give any warranty express or implied or make any
representation that the contents will be complete or accurate or up to
date. The accuracy of any instructions, formulae, and drug doses should
be independently verified with primary sources. The publisher shall not
be liable for any loss, actions, claims, proceedings, demand, or costs or
damages whatsoever or howsoever caused arising directly or indirectly in
connection with or arising out of the use of this material.
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013
Experimental Heat Transfer, 15:207–219, 2002
Copyright © 2002 Taylor & Francis
0891-6152 /02 $12.00 + .00
DOI: 10.1080/0891615029008264 9

A GENERAL HEAT TRANSFER CORRELATION FOR


IMPACTING WATER SPRAYS ON HIGH-TEMPERATURE
SURFACES

S. C. Yao
Department of Mechanical Engineering, Carnegie Mellon University, Pittsburgh,
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

Pennsylvania, USA

Timothy L. Cox
Alcoa Technical Center, Alcoa Center, Pennsylvania, USA

A method for the general correlation of heat transfer effectiveness for sprays impacting
vertically downward on a high-temperature surface has been developed . A dimensional
analysis showed that the mass velocity of the spray can be substituted for the droplet
velocity in the droplet Reynolds and Weber numbers, greatly improving the correlation with
the heat  ux data in the Ž lm boiling regime. The spray Reynolds number, deŽ ned as Res =
Gd/ m , and spray Weber number, Wes = G2 d/r j , were shown to correlate data from many
authors covering a wide range of spray parameters. This correlation supports the results
of previous parametric experiments, and is analogous to correlations developed for the
critical heat  ux of sprays and circular jets. Dependence of the heat transfer performance
on spray mass  ux and droplet diameter represents the in uence of the overall heat transfer
capacity of the spray as well as the contribution of droplet interactions. The Leidenfrost
temperature of the spray was also shown to be dependent on the spray Weber number.

The heat transfer of sprays impinging on high-temperature surfaces has been widely
studied. To date, however, there has not been a general correlation relating the heat transfer
to spray parameters which covers a wide range of spray operating values. Many articles
have presented equations that apply to the extent of the parameters for that particular
study [1–12], but these correlations generally fail outside their intended range. Further
hindering a broad correlation is the variety of presentation methods in the literature and
the lack of cross comparisons between different authors’ data.
Previous studies [1, 2, 5, 13–15] have shown a strong relationship between heat
transfer and the spray mass  ux. Also, there is evidence [1, 6, 7, 14–16] that droplet
diameter also in uences spray heat transfer performance. This implies that these parame-
ters are instrumental in the characterization of the spray heat transfer process. The effects
of droplet velocity have not been as clear [4, 15, 17, 18].
Recently, data were obtained [14] for sprays of large droplets with diameters rang-
ing from 3 to 25 mm, which had not been previously studied in the literature. It is of

Received 10 March 2001; accepted 12 September 2001.


Address correspondenc e to Prof. S. C. Yao, Department of Mechanical Engineering, Carnegie Mellon
University, 5000 Forbes Avenue, Pittsburgh, PA 15213, USA. E-mail: sy0d@andrew.cmu.edu

207
208 S. C. YAO AND T. L. COX

NOMENCLATURE
cp speciŽ c heat, J/kg K ½ density, kg/m3
d droplet diameter, m ¾ surface tension, N/m
G liquid mass  ux, kg/m2 s
h enthalpy, J/kg Subscripts
1hlg latent heat of vaporization, J/kg d droplet
n spray droplet density, m¡3 l property of liquid phase
q heat transfer rate, W L characteristic length
q 00 heat  ux, W/m2 Leidenfrost temperature at which minimum
Re droplet Reynolds number (D ½vd=¹) Ž lm boiling heat  ux occurs
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

T temperature, K s spray
v droplet velocity, m/s sat liquid saturation properties
V volume, m3 stream properties of a stream of drops
We droplet Weber number (D ½v2 d=¾ ) sub surface subcooling
" heat transfer effectiveness v property of vapor phase
¹ dynamic viscosity, kg/m s w conditions at target surface

interest to examine if this additional information allows the formulation of a general


correlation for spray heat transfer in the Ž lm boiling region that will apply to a large
range of spray operating conditions.
An attempt was made to correlate a limited set of heat transfer results, with some
success [19]. In the present article, the effects of liquid subcooling and variation of surface
temperature are considered in the database and new correlation. Comparison with other
predictive formulations and the explanation of the present model are provided. In addition,
the Leidenfrost temperatures are also shown to correlate over a wide range of spray
data.

DIMENSIONAL ANALYSIS
Traditionally, the most effective method of correlating experimental data has been
through the use of nondimensional groupings. To determine the groupings most appro-
priate to the nonwetting spray heat transfer process, a dimensional analysis was per-
formed.
Consider a subcooled droplet impacting a heated surface above the Leidenfrost
temperature (Figure 1). For a constant liquid and surface temperature, the most relevant
variables are G, v, d, q 00, ½, ¾ , ¹, and 1h. (Other variables such as surface roughness
[20], target material [4], and target dimensions have also been shown to in uence the
heat transfer to a lesser extent, but they will not be considered here.)
The liquid heat transfer capacity, 1h, represents the total amount of heat the liquid
is able to absorb, which is deŽ ned as

1h D hlg C cp;l .Tsat ¡ Tl / C cp;v .Tw ¡ Tsat / (1)

where Tsat is the liquid saturation temperature.


CORRELATION FOR IMPACTING SPRAYS ON SURFACES 209
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

Figure 1. Parameters for dimensional analysis of droplet impact


heat transfer.

Using the Buckingham pi method of nondimensionalization, Ž ve nondimensional


terms were obtained:

G=½v (2a)
¹=½vd (2b)

¾=½v 2 d (2c)

q 00=½v 3 (2d)

1h=v 2 (2e)

Reviewing this group, it is readily observed that the droplet Reynolds and Weber
numbers are generated, Eqs. .2b/ and .2c/. Additionally, combining the dependent-
variable term of q 00 in Eq. .2d/ with Eqs. .2a/ and .2e/ returns the spray heat transfer
effectiveness:

q 00
"D (3)
G[hlg C cp;l .Tsat ¡ Tl / C cp;v .Tw ¡ Tsat /]

The parameter " deŽ ned in Eq. (3) is referred to as the spray effectiveness. It represents
the amount of heat removed by the spray as related to the spray’s maximum heat transfer
potential, the sum of the latent heat and sensible heat capacity of the liquid. This term
is often used to represent a spray’s heat transfer efŽ ciency.
210 S. C. YAO AND T. L. COX
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

Figure 2. Comparison between effectiveness and .a/ droplet Reynolds number and .b/ droplet Weber
number.
CORRELATION FOR IMPACTING SPRAYS ON SURFACES 211

The effects of the droplet Weber and Reynolds numbers on effectiveness are shown
in Figure 2. Data in the graphs were collected from various authors [3, 6, 7, 13, 14] cov-
ering a wide range of conditions, including both monodisperse and polydisperse sprays.
The range of mass  uxes represented is 0.016–2.05 kg/m2 s, droplet velocities vary be-
tween 0.6 and 7.3 m/s, and droplet diameters of 0.13 to 25 mm are included. Although
there is much more data presented in the literature, many sources could not be included
with this study due to the omission of one or more spray parameters that are needed to
include them with the general correlation. Also, since this article is addressing sprays
and the interference between multiple droplets, results from single-droplet studies were
not included.
For monodisperse sprays, the actual droplet diameter could be used. However,
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

polydisperse sprays produce a spectrum of droplet sizes. In this case, the Sauter mean
diameter (d32 ) was used. This diameter represents a droplet whose volume-to-surface
area ratio is the same for that of the spray as a whole.
To promote comparability, all heat transfer data was taken at 400± C. Additionally,
the data of Shoji et al. [7] were corrected using the results of Deb and Yao [9] to account
for the geometry of their test, which eliminated the effects of secondary droplet impact.
In general, the effectiveness decreases with increasing We, with a jump at around
We D 80. This is the same Weber number value that Wachters and Westerling [21] relate
to a change in the dynamic behavior of a droplet as it impacts a nonwetting surface.

GENERAL CORRELATION
The Ž rst term of the dimensional analysis, Eq. .2a/, corresponds to the ratio of
droplet velocity to the spray’s mass velocity. The mass velocity of a spray represents the
liquid’s mass  ux—the rate of liquid impacting the surface. This can be interpreted by
visualizing a layer of water rising from the surface due to the impingement of the spray.
The rate of rise of the water level is G=½. Using this term, this characteristic velocity
can be substituted for droplet velocity in the droplet Reynolds and Weber numbers to
generate two new spray-related parameters,

Gd
Res D (4a)
¹

and

G2 d
Wes D (4b)
½¾

The modiŽ ed Weber and Reynolds numbers now represent the effects of the inertia of
the spray as a whole, rather than the individual droplets.
Figure 3a shows the relationship between the spray Reynolds number and the heat
transfer effectiveness. It can be seen that this parameter greatly improves the correlation
with the data. Interestingly, this spray Reynolds number is analogous to one Rohsenow
[22] used to generate his correlation for nucleate boiling data.
Even greater improvements are made when comparing effectiveness to the spray
Weber number (Figure 3b), effectively comparing performance for a wide range of spray
parameters.
212 S. C. YAO AND T. L. COX
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

Figure 3. Correlation of spray heat transfer effectiveness with .a/ spray Reynolds number and .b/ spray
Weber number.
CORRELATION FOR IMPACTING SPRAYS ON SURFACES 213

This correlation is expected to vary for different surface and liquid temperatures.
To include these effects, a nondimensional temperature parameter will be introduced.
Consider the characteristic temperatures of the process. The Ž rst is the subcooling of the
drop, 1Tsub D Tsat ¡ Tl , which represents the preboiling cooling potential of the liquid.
The superheat temperature difference is represented by 1Tsat D Tw ¡Tsat , providing
the driving potential for boiling. Combining these gives a total temperature difference of

1T D 1Tsub C 1Tsat D Tsat ¡ Tl C Tw ¡ Tsat D Tw ¡ Tl (5)

This is easily recognized as the temperature difference used in Newton’s law of cooling.
The characteristic temperature of the liquid for nondimensionalization is chosen to be
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

the liquid’s saturation temperature, Tsat .


As the superheat increases above the Leidenfrost temperature, it is expected that
effectiveness of the spray will decrease. The disintegration of the droplets becomes more
violent, allowing less time and contact on the surface. Therefore, the nondimensionalized
temperature difference will be introduced as

Tsat
(5a)
1Tsub C 1Tsat

by combining it with the Res and Wes .


The results of this modiŽ cation can be seen in Figures 4a and 4b. Including the
effects of temperature in the comparison has allowed for a general correlation of the data
for the range of surface superheats (1Tsat ) from 300 to 700± C.
Inclusion of higher-temperature data allows for the expansion of the database to
include additional industrial polydisperse sprays [2, 5, 23], expanding the range of surface
temperatures to 800± C, the mass  ux range to 50.5 kg/m2 s, and the droplet velocity to
20 m/s.
Correlations representing the least-squares Ž t between spray heat transfer effective-
ness and the spray Reynolds and Weber numbers are presented in Eqs. .6a/ and .6b/,
respectively. The spray Weber number, Wes , correlates with the data better than Res .

" D 2:5 ´ 10¡4 × [Res × Tsat =.1Tsub C 1Tsat /]¡1:05


(6a)
C 2:5 ´ 10¡2 × [Res × Tsat =.1Tsub C 1Tsat /]¡0:4

" D8´ 10¡7 × [Wes × Tsat =.1Tsub C 1Tsat /]¡0:62


(6b)
C 3:5 ´ 10¡3 × [Wes × Tsat =.1Tsub C 1Tsat /]¡0:2

The correlation with spray Weber number has an average error of 17% and a correlation
coefŽ cient (R 2 ) of 0.94 over the range 6´ 10¡10 < Wes < 3 ´ 10¡2 , whereas the average
error of Eq. .6a/ is 33% with an R 2 of 0.95 for 2 ´ 10¡3 < Res < 5 ´ 101 . These
correlations are applicable to mono- and polydisperse sprays. To account for the spatial
variability of a polydisperse spray, the mass  ux distribution of the spray is determined
and the correlation is applied locally in the region under the spray.
Dependence on spray Weber number re ects general spray behavior, in that droplet-
related phenomena are frequently Weber number dependent. Spray Weber number depen-
dence is also suggested in the results of parametric experiments [14]. These experiments
214 S. C. YAO AND T. L. COX
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

Figure 4. General correlation of spray heat transfer effectiveness with .a/ spray Reynolds number and
.b/ spray Weber number.
CORRELATION FOR IMPACTING SPRAYS ON SURFACES 215

Table 1. Prediction error of Eq. .6b/ and published models

Min Max Ave

Eq. .6b/ 0.1% 64.4% 17.3%


Ito 0.4% 204.9% 28.3%
Bolle 0.7% 500.4% 52.7%
Labeish 0.1% 411.2% 54.7%
Yao & Choi 2.7% 237.9% 96.2%
Mizikar 0.5% 385.7% 105.0%
Klinzing et al. 2.5% 546.7% 176.1%
Bernardin et al. 99.4% 453.3% 248.2%
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

showed that heat transfer effectiveness varied with the mass  ux as " / 1=G1=4 , and with
droplet diameter as " / 1=d 1=2 . Therefore, it would be expected that the mass  ux should
appear in a dimensionless correlation parameter to the second power of that of droplet
diameter. Also, terms analogous to the spray Weber number have been used to correlate
spray critical heat  ux [24] and the heat transfer effectiveness of circular jets [25].
At very low mass  uxes, some of the reported heat transfer data corresponded to
an effectiveness greater than 1.0. Although this by deŽ nition is not possible, it is most
likely due to the contribution of co- owing air, variations in surface characteristics, and
assumptions in the data reduction methods.

COMPARISON WITH OTHER MODELS


Several models have been published previously to predict the heat transfer of sprays
on high-temperature surfaces [1–6]. These models are mainly empirical in nature, and
were developed only for the range of spray parameters used in each author’s experiments.
Each independent model was compared with the overall database only in its intended
range of validity.
Table 1 shows the average error of each model in percent difference from the
experimental values. Equation .6b/ matches the data with a 17.3% average error. The
closest model from published literature is that of Ito [5] at 28.3%, primarily because most
of the data in this study that fell in the range of validity of their model were extracted
from their article.

INTERPRETATION OF Rs AND Wes


To understand why the spray Weber number correlates better than the droplet Weber
number, the in uence of mass  ux on the spray cooling process should be examined.
The mass  ux is related directly to the amount of heat removal capacity of the spray and,
along with the droplet diameter, is an indication of the interaction between the droplets
on the surface [1, 15].
Consider a stream of single droplets impacting an ideal surface (which maintains
a uniform wall temperature, and without droplet interaction). Each drop carries away a
Ž xed amount of heat, qd , depending on its temperature, size, and velocity. The amount
of heat transferred from a stream of droplets is proportional to the number of droplets
216 S. C. YAO AND T. L. COX

impacting the surface per unit time—the mass  ux. As the frequency of impactions
increases, the heat removal should increase linearly.

qd D q.T ; v; d/ (7a)
qstream D f × qd (7b)

This theory can be transposed to a monodisperse spray. Keeping v, d, and Tl


constant, the amount of heat removed is proportional to the number of droplets striking
the surface per unit time. Still neglecting droplet interaction and surface chilling, the heat
transfer should again increase linearly with droplet frequency, as depicted in Figure 5.
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

However, for a real spray there will be interaction between drops. This interaction
will be related to the density of the droplets on the surface and to the size of the droplets.
For droplets with constant velocity and diameter, the droplet density increases with the
spray mass  ux, increasing the probability of droplet interactions:

½¼ d 3 vn
G D ½8d vn D (8)
6

Although the droplet velocity does not appear directly in the spray Reynolds and
Weber numbers, it can be seen from Eq. (8) that the effects of droplet velocity are
included in the mass  ux term.
As droplets collide on the surface, the effective area and momentum of each drop
is reduced by competition with interfering drops, as noted by Bernardin et al. [1] and
Fujimoto et al. [15]. Consequently, as interaction increases, the effectiveness of the sprays
decreases.
At very high mass  uxes,  ooding of the surface occurs; that is, previous drops
do not have time to get out of the way of following ones. The new drops impact the
water on the surface Ž rst, decreasing their momentum and increasing their temperature
by mixing with the warmer spent drops. This phenomenon has been observed by Deb
and Yao [9].
Also, for a real surface, the increased heat  ux associated with an increase in mass
 ux causes chilling at the interface. The next droplet impacting the surface will encounter
a surface of lower temperature than the previous drop. As the droplet density increases,
the time available for the surface temperature to recover decreases. Since the heat transfer
in the nonwetting regime decreases with decreasing temperature, this sequential impacting
will also contribute to the decrease in effectiveness with increasing mass  ux.

Figure 5. Expected behavior of the heat


transfer with respect to spray mass  ux.
CORRELATION FOR IMPACTING SPRAYS ON SURFACES 217
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

Figure 6. Leidenfrost temperature variation with spray Weber number.

CORRELATION OF LEIDENFROST TEMPERATURE WITH Wes


Because increased chilling of the surface occurs with increased mass  ux, it would
be expected that the Leidenfrost temperature of a spray on a hot surface would also vary
with mass  ux. This effect has been noted in the literature [5, 14].
If the Leidenfrost temperatures from published spray heat transfer experiments are
compared with the spray Weber number (Figure 6), it can be seen that a relationship
exists between the two parameters, with the Leidenfrost temperature increasing with
increasing Wes .
This is due to the increasing surface chill and temperature gradient in the target
with the increased heat transfer of higher-Wes sprays. The liquid sees a temperature
much less than the bulk or average temperature of the target. Other parameters such as
surface roughness and thermal conductivity also in uence the Leidenfrost temperature,
but clearly there is a strong relationship with the spray Weber number.
The large scatter in the data in Figure 6 is due to the difŽ culty in selecting the
exact point of minimum heat  ux. As the Wes increases, the sharpness of the transition
point decreases. The data of Figure 6 suggests the relationship

TLeidenfrost D 1,400 × We0:13


s (9)

CONCLUSIONS
A dimensional analysis showed that the spray mass velocity, G=½, could be sub-
stituted for the droplet velocity in the traditional droplet Reynolds and Weber numbers,
greatly improving the correlation between those parameters and the spray heat transfer
218 S. C. YAO AND T. L. COX

effectiveness. These new spray-related parameters correlated heat transfer effectiveness


for a large range of spray operating parameters, spray types, and testing methods.
In general, the spray Weber number, because of its inclusion of the mass  ux,
re ects the behavior of real sprays more accurately than the traditional droplet Weber
number. Dependence of the spray heat transfer performance on the spray mass  ux and
droplet diameter represents the in uence of the overall heat transfer capacity of the spray
as well as the contribution of droplet interactions. This correlation re ects the results of
previous parametric studies showing a strong relationship between heat transfer and the
spray mass  ux.
The model presented in Eq. .6b/ demonstrates that the spray Weber number can be
used to predict spray heat transfer performance over a wide range of operating parameters.
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

The Leidenfrost temperature for a spray impacting a heated surface was shown to
be related to the spray Weber number as well. This is due to the increased surface chill
caused by the increased heat transfer at higher mass  uxes.
This study demonstrates that there is a strong relationship between the spray Weber
number and Ž lm boiling heat transfer and Leidenfrost temperature, and provides a foun-
dation on which to build a larger database and more detailed correlations. Future work
should address expanding the present database to include a larger range of liquid sub-
coolings and spray parameters (mass  ux, velocity, and droplet diameter). Additional
effort could also include a better understanding of the in uence of interaction and pool-
ing of liquid on the surface at high mass  ux, and veriŽ cation of the relationship be-
tween Leidenfrost temperature and spray Weber number including the effects of surface
conditions.

REFERENCES
1. J. D. Bernardin and I. Mudawar, Film Boiling Heat Transfer of Droplet Streams and Sprays,
Int. J. Heat Mass Transfer, vol. 40, no. 11, pp. 2579–2593, 1997.
2. L. Bolle and J. C. Moureau, Spray Cooling of Hot Surfaces, Multiphase Science and Technol-
ogy, Hemisphere, Washington, DC, 1982.
3. T. Ito, Y. Takata, M. M. M. Mousa, and H. Yoshikai, Studies on the Water Cooling of Hot
Surfaces (Experiment of Spray Cooling), Mem. Faculty Eng. Kyushu Univ., vol. 51, no. 2,
1991.
4. V. G. Labeish, Thermohydrodynamic Study of a Drop Impact against a Heated Surface, Exp.
Thermal Fluid Sci., vol. 8, pp. 181–194, 1994.
5. E. A. Mizikar, Spray Cooling Investigation for Continuous Casting of Billets and Blooms, Iron
and Steel Engineer, pp. 53–60, June 1970.
6. S. C. Yao and K. J. Choi, Heat Transfer Experiments of Mono-dispersed Vertically Impacting
Sprays, Int. J. Multiphase Flow, vol. 13, no. 5, pp. 639–648, 1987.
7. M. Shoji, T. Wakunaga, and K. Kodama, Heat Transfer from a Heated Surface to an Impinging
Subcooled Droplet (Heat Transfer Characteristics in the Non-wetting Regime), Trans Jpn. Soc.
Mech. Eng., vol. 50, no. 451, pp. 716–722, 1984.
8. S. Inada and W.-J. Yang, Film Boiling Heat Transfer for Saturated Drops Impinging on a
Heated Surface, Int. J. Heat Mass Transfer, vol. 37, no. 16, pp. 2588–2591, 1994.
9. S. Deb and S. C. Yao, Analysis on Film Boiling Heat Transfer of Impacting Sprays, Int. J.
Heat Mass Transfer, vol. 32, no. 11, pp. 2099–2112, 1989.
10. C. J. Hoogendorn and R. den Hond, Leidenfrost Temperature and Heat-Transfer CoefŽ cients
for Water Sprays Impinging on a Hot Surface, Proc. 5th Int. Heat Transfer Conf., vol. 4,
pp. 135–138, 1974.
CORRELATION FOR IMPACTING SPRAYS ON SURFACES 219

11. W. P. Klinzing, J. C. Rozzi, and I. Mudawar, Film and Transition Boiling Correlations for
Quenching of Hot Surfaces with Water Sprays, J. Heat Treating, vol. 9, no. 2, pp. 91–103,
1992.
12. I. Mudawar and W. S. Valentine, Determination of the Local Quench Curve for Spray-Cooled
Metallic Surfaces, J. Heat Treating, vol. 7, pp. 107–121, 1989.
13. K. J. Choi and S. C. Yao, Mechanisms of Film Boiling Heat Transfer of Normally Impacting
Spray, Int. J. Heat Mass Transfer, vol. 30, no. 2, pp. 311–318, 1987.
14. T. L. Cox and S. C Yao, Heat Transfer of Sprays of Large Water Drops Impacting on High
Temperature Surfaces, J. Heat Transfer, vol. 121, no. 2, pp. 446–450, 1999.
15. H. Fujimoto, N. Hatta, H. Asakawa, and T. Hashimoto, Predictable Modelling of Heat Transfer
CoefŽ cient between Spraying Water and a Hot Surface above the Leidenfrost Temperature, ISIJ
Downloaded by [University of California Santa Barbara] at 05:45 28 June 2013

Int., vol. 37, no. 5, pp. 492–497, 1997.


16. D. M. Parke and J. L. L. Baker, Temperature Effects of Cooling Work Rolls, AISE Report.
17. C. O. Pederson, An Experimental Study of the Dynamic Behavior and Heat Transfer Charac-
teristics of Water Drops Impinging upon a Heated Surface, Int. J. Heat Mass Transfer, vol. 13,
pp. 369–381, 1970.
18. M. H. Shi, T. C. Bai, and J. Yu, Dynamic Behavior and Heat Transfer of a Liquid Droplet
Impinging on a Solid Surface, Exp. Thermal Fluid Sci., vol. 6, pp. 202–207, 1993.
19. T. L. Cox and S. C. Yao, Heat Transfer Experiments and Correlation for Water Sprays on High
Temperature Surfaces, Heat Transfer 1998, vol. 3, pp. 201–206, 1998.
20. J. D. Bernardin, C. J. Stebbins, and I. Mudawar, Effect of Surface Roughness on Water Droplet
Impact History and Heat Transfer Regimes, Int. J. Heat Mass Transfer, vol. 40 no. 1, pp. 73–88,
1997.
21. L. H. J. Wachters and N. A. J. Westerling, The Heat Transfer from a Hot Wall to Impinging
Water Drops in the Spheroidal State, Chem. Eng. Sci., vol. 21, pp. 1047–1056, 1966.
22. W. M. Rohsenow, A Method of Correlating Heat-Transfer Data for Surface Boiling of Liquids,
Trans. ASME, pp. 969–976, August 1952.
23. H. Yu, Unpublished data, Alcoa Technical Center, 1997.
24. K. A. Estes and I. Mudawar, Correlation of Sauter Mean Diameter and Critical Heat Flux for
Spray Cooling of Small Surfaces, Int. J. Heat Mass Transfer, vol. 38, no. 16, pp. 2985–2996,
1995.
25. Y. Katto and S. Yokoya, Critical Heat Flux on a Disk Heater Cooled by a Circular Jet of
Saturated Liquid Impinging at the Center, Int. J. Heat Mass Transfer, vol. 31, no. 2, pp. 219–
227, 1988.

You might also like