You are on page 1of 21

Rewetting and boiling in jet impingement on high temperature steel surface

H. Leocadio, C. W. M. van der Geld, and J. C. Passos

Citation: Physics of Fluids 30, 122102 (2018); doi: 10.1063/1.5054870


View online: https://doi.org/10.1063/1.5054870
View Table of Contents: http://aip.scitation.org/toc/phf/30/12
Published by the American Institute of Physics
PHYSICS OF FLUIDS 30, 122102 (2018)

Rewetting and boiling in jet impingement on high temperature


steel surface
H. Leocadio,1 C. W. M. van der Geld,2 and J. C. Passos3
1R & D Center, Usiminas Steel, Ipatinga, Brazil
2 Department of Chemical Engineering and Chemistry, Eindhoven University of Technology,
Eindhoven, The Netherlands
3 Department of Mechanical Engineering, Federal University of Santa Catarina, Santa Catarina, Brazil

(Received 4 September 2018; accepted 9 November 2018; published online 3 December 2018)

Interface topologies and boiling phenomena are observed within the water jet impingement zone
during quenching of a high temperature (300 ◦ C–900 ◦ C) steel plate by direct optical observations.
Stable film boiling may occur, but surface asperities may easily penetrate the vapor film and interact
with the flowing liquid. By cooling down more rapidly than the remaining solid surface, such asperities
act as a kind of micro-fin. Rather surprisingly, non-coalescing bubbles on top of the thin vapor film
have been observed, probably formed on top of penetrating surface asperities. After establishing
contact between water and solid, so after rewetting, an intense bubble activity is normally seen of
bubbles that are footed on a dry plate area with a contact line reflecting the light-emitting diode lights
and condensing at later times. These are vapor bubbles. A high surface roughness promotes rewetting
that may be initiated at several places simultaneously. For an initial plate temperature of 300 ◦ C and a
total height roughness of 5 µm, rewetting took place without the occurrence of a vapor film; surfaces at
or above 450 ◦ C exhibited vapor film formation, even for a high jet subcooling of 80 ◦ C. Temperature
and time of rewetting are strongly affected by the initial surface temperature and jet subcooling and
less by the jet velocity. New correlations are provided to predict these dependencies. Published by
AIP Publishing. https://doi.org/10.1063/1.5054870

I. INTRODUCTION of flow regimes are anticipated. According to Karwa et al.,2


it is not known whether bubbles exist in the wetted region.
Water jet impingement quenching has high cooling poten-
Knowledge of interfacial topologies and the occurrence of
tial and is used as an effective means of cooling in many
bubbles is found to be critical for the development of heat
industrial applications requiring rapid cooling. Examples are
transfer models for jet impingement quenching, but due to
emergency core cooling in nuclear power plants and power
the harsh conditions prevailing in the quenching zone and the
dissipation in the electro-mechanical devices. Steel industries
resulting near-inaccessibility, no direct optical observations
widely employ the process for accurate temperature control to
could be carried out until now. The only optical observations
improve the microstructure and to ensure proper mechanical
that were performed were from the outside of the jet look-
and metallurgical properties. The required surface tempera-
ing in Refs. 2–5. In the present study, interfacial topologies
tures, heat fluxes, and cooling rates are typically very large
are directly observed in the course of time and linked to the
(above 6 MW/m2 and above 100 ◦ C/s).
corresponding temperatures and heat fluxes occurring at the
Interface topologies and flow patterns can be quite com-
plate.
plex in boiling from high temperature surfaces.1 The same is
Quite some concepts related to quench cooling have been
expected during quench cooling of steel plates at 900 ◦ C, at
defined in the literature, sometimes with different names for
ambient pressure, but flow regimes have never been directly
the same concept. To be clear about the names employed in
observed in these circumstances, yet. One of the reasons is
this paper, the following list of definitions is given:
the difficulty in accessing the rather confined liquid flow area
where large temperature gradients occur and where various 1. Quench or quenching: the process of transient, rapid
boiling processes are expected to occur.2 In hot strip rolling cooling of a hot plate. In metallurgy, this method is used
mills, the steel strip has a temperature of about 800–950 ◦ C to obtain specific microstructure and mechanical prop-
after the last finish stand mill and is then cooled at controlled erties. In loss of coolant accidents, it is used to rapidly
rates by arrays of impinging water jets (planar or circular) cool down reactor elements.
placed at a row along the length of the run-out table, while the 2. Rewetting: the process of establishing direct liquid con-
plate is moving at speeds up to 20 m/s. As initially the strip tact with the plate when a vapor layer is initially cov-
surface temperature is far above saturation temperature, an ering the plate. A contact line between liquid, vapor,
insulating vapor blanket covers the solid surface (film boiling) and solid is established which is often called rewetting
in this stage. During cooling down to a temperature that allows front. This contact line can be moving. In rewetting,
a stable metal-liquid contact (rewetting), various other types transition from film boiling through a vapor layer to

1070-6631/2018/30(12)/122102/20/$30.00 30, 122102-1 Published by AIP Publishing.


122102-2 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

a heat transfer process between a solid and a liquid parameters are carefully examined and quantified in the present
occurs. study.
3. Stagnation zone: the area on the flat plate just under- One of these parameters, the rewetting temperature, Trw ,
neath the approximately cylindrical water jet, with a was found to vary in a wide range. In some water jet
diameter equal to the smallest diameter occurring in the quenching10 and reflooding11 studies, rewetting was found
jet. to occur only at solid surface temperatures below 300 ◦ C
4. Partial wetting: gas-liquid interface topology stage with and 308 ◦ C, respectively. Also, analytical studies12,13 pre-
one or more contact lines occurring in the stagnation dicted rewetting interface temperatures ranging from 303 ◦ C to
zone. 325 ◦ C. Nonetheless, a number of experimental studies14–20
5. Leidenfrost temperature: the wall temperature at which have reported rewetting taking place at surface temperature
a drop that was floating on a vapor cushion collapses above 700 ◦ C which even exceeds the critical point of water.
and touches the surface. Karwa and Stephan,20 for example, reported a rewetting tem-
6. Rewetting temperature, Trw : the temperature at which perature of 650 ◦ C for a water jet subcooling of 75 ◦ C, with a
stable solid-liquid contact is established during quench corresponding heat flux above 6 MW/m2 . Pandey and Basu21
cooling; this temperature marks the onset of rapid plate and Lapenna22 claimed that rewetting takes place always if the
surface cooling. The rewetting temperature is the aver- water jet subcooling exceeds 40 ◦ C, regardless of the initial
age temperature at the top of the plate in the stagnation plate surface temperature. Although Hall et al.1 claimed that
zone. This zone contains not only wetted locations but the high rewetting temperatures reported by Yazdani et al.,23
also parts in contact with vapor, as will be seen in this Pandey and Basu,21 and Filipovic et al.8 should be consid-
paper. The rewetting temperature is sometimes named ered as physically impossible, more recent publications21–28
quenching temperature. Alternative definition would be have shed new light on the rewetting and Leidenfrost temper-
the minimum wall temperature necessary for the exis- atures that help to understand the wide range of Trw -values
tence of a stable vapor layer blanketing the surface of found. The mechanism of boiling depends on a wide vari-
the plate, and then the rewetting temperature may also ety of physical process and material conditions,21–23 amongst
be called minimum film boiling temperature. If the liq- which the presence of nanoparticles is one.24 Such effects also
uid has a homogeneous saturation temperature and if occur during film boiling.25–27 Kim et al.28 performed a most
flow is not affecting the gas-liquid interface, the mini- interesting study of the effects of surface roughness height,
mum film boiling temperature is equal to the Leidenfrost wettability, and nanoporosity on wetting. A nanoporous silicon
temperature. oxide surface made of silicon oxide nanoparticles was cooled
7. Rewetting delay: the time between the start of cooling by water droplets, and the Leidenfrost temperature could attain
and the onset of rewetting. values as high as 453 ◦ C on this surface, which is higher than
8. Thermodynamic limit of water superheating, Ttls : the the critical temperature of water. The surface roughness has
maximum water superheating that is possible at a given a big influence and explains the spread in experimental find-
pressure. At atmospheric pressure, the corresponding ings of the rewetting temperature. Also in the present study,
maximum temperature of the solid-liquid interface that the effects of surface asperities at the hot solid surface on
is possible without instantaneous vaporization of the rewetting will be examined. Rewetting phenomena will be
liquid is 325 ◦ C.6 Instantaneous vaporization not nec- observed directly within the jet impingement zone. The rewet-
essarily occurs at temperatures below 325 ◦ C.7 At solid- ting temperature will be coupled to interface topologies by
liquid interface temperatures exceeding 325 ◦ C, the direct, optical observation.
liquid is repelled from the hot surface and wetting or Under equal material plate surface conditions, process
rewetting cannot occur.8,9 Another phase limiting con- parameters may affect the rewetting temperature and the rewet-
dition for water is of course the critical temperature of ting delay. Wang et al.4 concluded for water jet quenching
374 ◦ C at a pressure of 22 MPa. experiments of a hot steel plate (200–900 ◦ C), jet subcool-
9. Plate superheating and jet subcooling: The plate super- ing 13–87 ◦ C, that the rewetting temperature is profoundly
heating is the difference between the plate temperature affected by the initial plate temperature, Ti, and almost unaf-
and the saturation temperature, T sat , at the prevailing fected by velocity and temperature of the jet. Takrouri et al.5
system pressure, which is the ambient pressure in our quenched horizontal tubes of 380–780 ◦ C with water to find
measurements. Similarly, the jet subcooling is defined that Trw is not affected by Ti but is strongly dependent on
as the difference between T sat and the temperature of jet subcooling ( jet subcooling in the range 20–75 ◦ C). Nobari
the impinging liquid jet. et al.32 proposed an empirical correlation for planar and cir-
10. Boiling curve: a plot of the heat flux at a place at the top cular water jets based on only the jet subcooling. Piggott
surface of the plate, in this study the stagnation zone, et al.33 and Mozumder et al.34 found the rewetting delay, trw , to
versus plate superheating. be a strong function of Ti and the jet velocity for jet subcool-
ings less than 30 ◦ C, and not occurring beyond. According
Efficient cooling control at elevated cooling rates starts to Agrawal et al.,18 the jet velocity has no influence on the
at the onset of rewetting. Accurate knowledge of the rewet- rewetting delay time but does affect Trw .
ting temperature and the rewetting delay is therefore essential In the present experimental study of water jet impinge-
to improve cooling models and to predict the cooling rate ment on a high temperature steel plate, high-speed imaging
in the so-called ultra-fast cooling technology.3–5 Rewetting (up to 20 kfps) is used to measure the interfaces within the
122102-3 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

jet impingement zone from above, with an optical path that until the desired temperature was reached everywhere. The
goes all through the impingement jet. The plate was pre- water jet subcoolings were 80 ◦ C, 50 ◦ C, and 30 ◦ C with jet
heated at 300 ◦ C–900 ◦ C and quenched by an 8 mm-impinging velocities of 1 and 3 m/s. The inner diameter of the nozzle is
circular water jet at temperatures ranging from 20 ◦ C 9.7 mm, and the nozzle is positioned centrally above the test
to 70 ◦ C, with jet velocities of 1 and 3 m/s. Rewetting is plate surface. The water flow rate is set by a needle valve. A
observed directly and coupled to the solid surface tempera- high-speed camera, #2 in Fig. 1(a), captures images from the
ture that is reconstructed by solving the inverse heat problem outside of the jet at 300 fps.
for temperature histories measured in the plate. The effects The test rig described so far is fairly standard in quench
of subcooling, jet velocity, and initial surface temperature on cooling. The one thing that discriminates our test rig from
the rewetting temperature and delay have been measured in those of other researchers is the possibility to optically observe
this way. Empirical correlations for all dependencies will be interfacial phenomena through the liquid jet on the hot surface
proposed. of the plate. To this end, a borescope is mounted on a box that
Most interestingly, two types of bubbles have been found is fully filled with liquid. The borescope is connected to the
to occur, one of them even simultaneously with the initial vapor high-speed video camera #1 and is positioned centrally above
film covering the solid surface. In order to unravel the nature the hole of the nozzle exit in a way that does not disturb the
of these bubbles, a step-by-step approach will be followed in water jet stability. In this way, recordings through the water
which first the characteristics of vapor bubbles in our mea- jet yield top views from the jet impingement zone at a rate
surements will be assessed. Next, it will be shown that the of 5–20 kfps. A series of light-emitting diode (LED) lights
other types of bubbles do not exhibit these features and fol- surrounding the nozzle exit provide sufficient illumination for
low quite different trajectories during moving above the hot this purpose.
plate. A cylindrical coordinate system is defined in which r is the
radial distance to the center of the jet and y is the height above
the surface of the plate. The y-direction is given by −g/|g| if g
II. EXPERIMENTAL SETUP AND TEST PROCEDURES denotes the gravity vector.
A. Experimental setup with borescope
B. Test plate
The main components of the experimental setup are
schematized in Fig. 1(a). Normal tap-water is stored in an Figure 1(b) shows a cross section of a jet of 8 mm width
open head tank in which the water temperature is controlled and a test plate of 50 × 50 × 10 mm3 , with 10 mm being the
by using a thermostat controller. Water is pumped from the thickness of the plate. The impinging jet diameter is similar
collecting reservoir to the head tank. An overflow tube kept to those used in actual hot strip mill cooling. The test plate
the water level in the head tank constant, which ascertains a material is made of traditional austenitic stainless steel type
constant pressure and a stable water jet at the nozzle exit. The 304, and this steel was chosen for two reasons. First, because
water was heated up by using an electrical immersion heater of its high chromium content (18%-20%) which ensures good
and recirculated through head tank, water box, and reservoir resistance against high temperature oxidation by the formation

FIG. 1. (a) Schematic of the test rig and (b) schematic of cross section of a jet impinging on test plate with thermocouples TC1, TC2, and TC3 located at a line,
at various radial positions in a single (r, y)-plane in a cylindrical coordinate system with axial coordinate y and radial coordinate r.
122102-4 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

of a protective chromium oxide (Cr2 O3 ) film.35 This film is TABLE II. Hydrodynamic parameters at the stagnation point (r = 0, y = 0).
formed prior to the quenching tests when 304 stainless steel
Qn (`/min) dn (mm) dj (mm) Vj (m/s) Pj (kPa) Tsat (◦ C)
(SS) is heated in air to elevated temperatures.35 Second, to
avoid a notable amount of heat generation, caused by phase 3 9.7 8 1 102 100
transformations inside the steel, as is common in carbon steel. 9 9.7 8 3 106 101
On the sides, 25 mm thick thermal insulation with low thermal
conductivity of 0.11 W/(m ◦ C) at 800 ◦ C is mounted.
Three holes of 1.1 mm width and 9 mm depth were drilled container. The jet velocity at the water exit follows easily from
and checked with calipers. Thermocouples TC1, TC2, and TC3 Vn = 4∗ Qn /(π dn 2 ). The uncertaintyqof Vn was δVn /Vn = ±3%.
are located at a line at radial positions of 0, 9, and 18 mm
The jet velocity follows from Vj = V2n + 2gH, where H is the
from the center of the plate, respectively, and at 1 mm below
height of the nozzle exit to the plate and g is the acceleration
the top surface. The tip of the thermocouple is placed 1 mm
of gravity. The jet diameter is then given by dj = dn Vn /Vj .
p
below the surface. The actual location of the sensing part of
The stagnation pressure is the sum of the ambient pressure
the thermocouple is determined by cutting the thermocouple
of 101 kPa and the part related to the impinging jet velocity:
after its use in the experiments and by measuring the distance
ρVj 2 /2.
to the tip of the thermocouple.
Heat transfer in jet impingement cooling of steel plates is
Grounded thermocouples type-K, 1 mm in sheath diam-
strongly dependent on how water flows over the surface, which
eter made of 304 SS (the same as the test plate), are used.
will be seen to affect the dynamics of the rewetting front. For
High temperature thermal paste with a conductivity of 70
this reason, jet impingement hydrodynamics on an unheated
W/(m ◦ C) was inserted into the hole to insure good thermal
flat surface are briefly described. As the circular free-surface
thermocouple-plate contact. The thermal properties of the 304
jet impinges onto the surface, the flow is spread out radially
SS36 are shown in Table I.
outward. The pressure is maximum at the stagnation point
The test plate was heated by using a manually operated
(r = 0) and decreases monotonically to the ambient pressure at
electrical oven (3 kW) placed over the plate in the stationary
about r/dj ≈ 1.28.29
quenching position underneath the borescope. The data acqui-
Based on studies of Stevens and Webb,37–39 profiles of the
sition system simultaneously triggers the automatic valve, the
normalized radial velocity, Vr /Vj , across the liquid layer are
high-speed camera, and LED lights and acquires and stores
given for several radial positions in Fig. 2. Details are given in
the data. The temperature histories of the three thermocouples
the Appendix. The velocity profiles actually occurring during
were measured at a rate of 50 Hz. The quenching tests were
quenching are of course different near the wall when a vapor
carried out at the initial plate temperature, Ti , of 300, 450,
film and bubbles are formed. Figure 2, nevertheless, indicates
600, 750, and 900 ◦ C. These are the temperatures which are
the velocity field that can be expected further away from the
commonly used in hot strip mill cooling.
wall and close to the center of the jet.
Before each experiment, the impinging surface was
TC1 is located at the stagnation point (r/dj = 0) where the
sanded with 320 grit sandpaper and cleaned with acetone. The
radial velocity is zero, while on TC2-position (r/dj = 1.12),
measured arithmetic mean roughness, Ra, at the impingement
the radial velocity can be about 20% higher than Vj near the
zone surface and after a quenching test was found to be in
surface (y ≈ 0.2 mm). On TC3-position (r/dj = 2.25), the radial
the range 0.11–0.18 µm, while for a new plate, it was 0.100
velocity is less than that at TC2-position, but it is about 10%
± 0.005 µm. The effect of the roughness on the rewetting in
larger than Vj , still.
the experiments is extensively discussed in Sec. III D.

C. Hydrodynamic parameters in jet impingement D. Uncertainty analysis

Some hydrodynamic parameters are relevant in a heat The primary parameters are rm , ym , t, Tm , Tj , and k that,
transfer analysis of jet impingement, such as the impinging respectively, represent the radial coordinate of thermocouples
jet velocity, Vj , jet diameter just before impingement, dj , and numbered m as measured from the center of the plate, the axial
stagnation pressure in the impingement zone, Pj . These param- coordinate of thermocouples measured from the bottom of the
eters are listed in Table II. The inner diameter of the nozzle plate, measured time, measured temperature, water jet tem-
of the water exit, dn , was 9.7 mm, and the volume flow rate perature, and thermal conductivity. The measurement errors
of water exiting, Qn , was measured by collecting water in a of these parameters and propagation of these uncertainties to
the calculated surface heat flux, qs , are presented in Table III.
All uncertainties (measured and calculated) in Table III are
TABLE I. Thermophysical properties of the 304 SS.
within 95% limits and are analyzed according to criteria sug-
T (◦ C) cp (J/kg K) ρ (kg/m3 ) k (W/m K) gested by Taylor.40 These criteria take errors in measurement,
calibration, machining, and measuring devices into account.
27 447 7900 15.2
The uncertainty in Vj is 3%.
127 515 7859 16.6
For the temperature measured internally in the plate, Tm ,
327 557 7774 19.8
527 582 7685 22.6
sheathed grounded K-type thermocouples have been used that
727 611 7582 25.4 provided a calibration accuracy of ±1.1 ◦ C or ±0.4% of reading
927 640 7521 28.0 (ASTM E230), whichever is greater, with uncertainties within
99% limits. As the test temperatures were 300–900 ◦ C, the
122102-5 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

FIG. 2. Profiles of the normalized


radial velocity, Vr /Vj , in the liquid layer
depth at various radial positions.

relative value of ±0.4% produces a greater uncertainty in mea- For this reason, the calculated relative uncertainty of the sur-
sured temperatures than ±1.1 ◦ C and was therefore selected for face heat flux at the top of the plate in the impingement zone
the uncertainty analysis. The data acquisition system produces is ±5%.
a mean error of ±0.17% within 99% of confidence level. Com- The plate is heated and covered by a heater until the very
bining the uncertainties of thermocouple and data acquisition last moment prior to the measurements. Directly after with-
system, the total uncertainty of the temperature measurement drawal of the cover, the top surface cools at a rate of 2 ◦ C/s.
system was ±0.43% within 99% limits or ±0.30% within 95% After touchdown of the jet, the cooling rate is about 400 ◦ C/s,
limits. which implies that the contribution by natural convection is
Although uncertainties in the thermocouple depths were less than 0.5% at times relevant to our analysis.
small, the effect of these measurements on calculated heat
fluxes and surface temperatures is significant. Thus, special E. Inverse heat conduction analysis
care has been taken on the real temperature reading position in
A commercial inverse heat conduction program,
the plate because an error in the reading position of, for exam-
INTEMP, was used to predict the heat flux and temperature
ple, 0.6 mm in depth would result in an error of 60% in the
distribution along the quenching surface from temperature his-
maximum heat flux on the quenching surface. The top sheath
tories measured with thermocouples inserted in the test plate.
thickness where the measuring junction is grounded inside
INTEMP has been applied successfully in several quenching
the thermocouple shielding may vary from 0.1 to 1.0 mm,
studies,14,41–44 and it allows the thermal properties of mate-
depending on the manufacturer. To ensure a high accuracy
rial to vary with temperature in each finite element. This
in the heat flux, the thermocouples were opened after the
is an important characteristic because it avoids errors up to
tests and the top sheath thickness was measured. The mea-
about 20% in estimating the surface heat flux for the fast
sured temperature histories were all combined to determine
and big temperature changes occurring in quench cooling.13
the accuracy of the heat flux in the reconstruction that will
Temperature-dependent thermal properties of 304 SS, shown
be described in Sec. II E. The heat diffusion of the material
in Table I, were used in the present analysis.
around the measuring junction is the same as that of the steel
INTEMP uses the dynamic programming method to solve
plate and the distance of the measuring junction to the surface
the nonlinear inverse heat conduction problem with the finite
is precisely known by cutting the couple after the experiments.
element method. It uses the Crank–Nicolson formulation to
solve the heat conduction model, together with a regulariza-
TABLE III. Uncertainties. tion term used to minimize the error in estimated fluxes. The
temperature boundary condition at bounding surfaces other
Quantity Uncertainty than the top of the plate is prescribed. The surface heat fluxes
rm (mm) δrm = ±0.05 assumed at the top of the plate are subsequently modified
t (s) δt = ±0.01 by a nonlinear optimization technique to minimize the error
k (%) δk/k = ±6 between the measured temperatures and the predictions. Con-
Measured
ym (mm) δym = ±0.05 vergence is reached once the heat flux reaches the optimum
Tm (%) δTm /Tm = ±0.3 value. The methodologies of modeling and computations were
Tj (◦ C) δTj = ±2 detailed by Trujillo and Busby.45
qs (%) δqs /qs = ±5 Calculated A 2D axisymmetric finite element model was used for
the analysis of the data of the present study. The model had
122102-6 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

this milky region to violent bubble nucleation and transport


processes.
A small darker zone can be observed at t = 0.27 s under
the jet, indicating that the solid-liquid contact (rewetting) has
taken place and that this region is wetted. Within this region,
the regime changes from film to nucleate boiling. Outside this
small region, the surface is dry and a vapor blanket (milky
region) occurs. After rewetting has started, the wetted region
grows fast moving radially outward (t = 0.38 s). At the outer
FIG. 3. 2D axisymmetric finite element model divided into three unknown periphery of this region, light is reflected at the curved vapor-
heat flux zones at the surface on top, where coordinate y is zero. The positions
of the measuring thermocouples are indicated by red dots.
liquid interface near the contact line, which results in a white
narrow band that demarcates the frontier between the wetted
zone and dry zone, named rewetting front.
25 mm radius and 10 mm thickness with 600 and 1000 The free-surface on top of the liquid flow appears smooth
quadratic elements of 0.5 × 0.5 and 0.5 × 0.2 mm2 , respec- and shiny, without visible evidence of boiling activity within
tively, with 4-nodes per element as shown in Fig. 3. The the wetted region (t = 0.44 s). Even though this visible absence
small elements of 0.5 × 0.2 mm2 filled the first 4 mm from of bubbles was reported in other studies,1,20,14 top view images
the top surface with thinner elements in the vertical direc- from within the jet will, in the present study, in Sec. III C,
tion because of the faster response that is needed close to the clearly show vigorous boiling activity. Outside of the wetted
quenching surface. The faces of the plate without jet impinge- region, the outgoing water is deflected upward from the surface
ment were considered adiabatic. The impingement surface was in the form of a thin sheet which eventually breaks up due to
divided into three unknown heat flux zones: r1 = 0–5 mm surface tension. This deflection, a kind of spattering away, was
(Zone 1), r2 = 5–14 mm (Zone 2), and r3 = 14–25 mm also observed in other jet quenching experiments1,20 and was
(Zone 3). The measured temperature histories are given as attributed to vigorous vapor production at the rewetting front
the input at the nodes corresponding to the positions of ther- location.
mocouples TC1, TC2, and TC3 located at 1 mm from the The rewetting front velocity, Vrw , decreases as the wetted
top surface at the radial position of r = 0, 9, and 18 mm, zone grows radially outwards. The same behavior was also
respectively. observed in our previous study14 and in other studies.46,20,19
The performance of this model was validated with the Note that the wetted zone diameter enlarges from about
aid of another commercial finite element analysis software, 16 mm (t = 0.38 s) to 44 mm (t = 0.68 s) in just 0.30 s,
ANSYS. For some typical test cases, the surface temperatures i.e., a radial advance of 14 mm. After a long period of 1.52 s
calculated by the ANSYS model were compared to those pre- (t = 2.20 s), the wetted zone diameter becomes about 46 mm,
dicted by the INTEMP model. The mean disagreement found i.e., a radial advance only of 1 mm. This sudden drop in velocity
between these results was 0.01%. is explained from the heating and thinning (Fig. 2) of the water
film while it moves radially outwards: it loses its ability to con-
dense vapor at rewetting the front edge. The free-liquid layer
III. RESULTS AND DISCUSSION thickness, hj , reduces from 1.6 to 0.6 mm at radial positions
The organization of this section is as follows. First, obser- 9 and 18 mm, respectively (Sec. II C). Times exceeding 2.2 s
vations of the classical type, from outside the jet, are presented will be discussed more fully in Sec. III C but in these experi-
and analyzed in Secs. III A and III B. Next, the novel obser- ments correspond to nucleate boiling occurring at all places.
vations of interface topologies, including those occurring in B. Typical cooling curves
boiling and rewetting, via top view images within in the jet
impingement zone, are presented in Sec. III C. Sections III D Figure 5(a) shows temperature histories measured by
and III E elucidate the mechanisms causing rewetting and using thermocouples located h = 1 mm under the top plate
quantify the related temperatures and delay times. at positions r = 0, 9, and 18 mm, for Ti = 900 ◦ C, ∆Tsub
= 80 ◦ C, and Vj = 1 m/s. Time at the onset of cooling by jet
impingement is set to zero. At this time, all the cooling curves
A. Visual observations of the jet quenching
coincide. Only low heat losses occur prior to this time, mainly
experiment from outside the jet
by radiation and convection to the ambient; the resulting rate
The sequence of snapshots of a jet quenching experiment of temperature decrease amounts to 2 ◦ C/s.
for Ti = 900 ◦ C (896 ◦ C to be precise, henceforth indicated About 0.1 s after the jet hits the surface, a sharp drop
with 900 ◦ C) and Tj = 20 ◦ C in Fig. 4 shows the growth of the in temperature is observed at all measurement positions. This
wetted region. The plate surface initially at 900 ◦ C becomes delay is due to the measuring position being 1 mm below the
covered by an insulating vapor blanket (film boiling) just after top surface, and the value of the corresponding Fourier number,
the jet has impinged on its surface (t = 0.06 s). The liquid moves Fo = k.ρ.t/(cp · L2 ), being about 0.6 confirms this. Here cp , k,
smoothly over the vapor blanket without splashing. This area ρ, t, and L are specific heat (J/kg K), conductivity (W/m K),
with the vapor film appears as a blurry whitish background, a time (s), and the characteristic length (m), respectively. The
“milky region,” from the outside for a reason that is not clear thermocouple distance of 1 mm below the top surface is the
“a priori.” Observations of Sec. III C are necessary to couple characteristic length here.
122102-7 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

FIG. 4. Example of history of growth of the wetted


region, Ti = 900 ◦ C and Tj = 20 ◦ C. The position of
the third thermocouple is at r/dj = 2.25, which is about at
the position of the rewetting front at t = 0.44 s.

The temperature drops in a less pronounced way at Figure 5(b) shows the radial distribution of surface heat
r = 9 mm than at r = 0 due to its greater radial distance from flux as a function of time. Initially, an insulating vapor blan-
the stagnation point, where the cooling intensity is larger. The ket covers the surface and the heat flux is low. The wetted
smallest cooling rate occurs at r = 18 mm when the surface zone increases with the advancement of the rewetting front, as
is not wetted, dry, and cooling occurs by radiation and vapor observed in Sec. III A. After rewetting front arrival, the heat
convection mainly. At about 1.2 s at this surface location, a flux increases up to reaching its maximum, qmax . Once the
remarkable increase in the cooling rate occurs when this spot maximum heat flux is reached, all heat flux curves decrease
becomes wetted by the arrival of the rewetting front. with about the same slope. The time when heat flux is at

FIG. 5. Measured cooling curves for


Ti = 900 ◦ C, ∆Tsub = 80 ◦ C, and Vj
= 1 m/s, which are the same conditions
as in Fig. 4: (a) measured temperature
histories at r = 0, 9, and 18 mm and at
1 mm below top surface and (b) histories
of surface heat flux above the measuring
points.
122102-8 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

maximum marks the time when nucleate boiling occurs at artist impressions based on recordings by using the side-view
all locations of the plate surface. After that time, all heat camera, HS-camera 2 in Fig. 1. The top views are, as far as
flux curves decrease with about the same slope because the the authors know, the first recordings ever of nucleate boiling
same boiling regime occurs at each and every radial position, during quench cooling of a horizontal plate.
yielding about the same heat transfer coefficient anywhere. Surprisingly, the hot surface becomes completely wetted
The heat flux peaks occur at different times at the radial as the jet touches down on it. No vapor filmed is formed, despite
positions of 0, 9, and 18 mm due to the delay in rewetting front the superheating of 200 ◦ C. It is important to note that the
arrival. Section III E will couple temperature and heat flux to side-view camera did not show a milky region such as the one
interface topologies also at times far after rewetting. shown in Fig. 4 which makes clear that no vapor film is formed
in this case with Ti = 300 ◦ C. The non-existence of a vapor
C. Film boiling, transition, rewetting, film at the plate is easily explained because the surface tem-
and nucleate boiling perature, 300 ◦ C, is below the thermodynamic limit of water
The topology of vapor-liquid-solid interfaces during superheating, Ttls , of 325 ◦ C;6 at temperatures exceeding Ttls ,
quench cooling by a circular water jet depends strongly on the a fluid undergoes instantaneous vaporization at atmospheric
initial surface temperature, Ti , jet subcooling, ∆Tsub , imping- pressure.7
ing jet velocity, Vj , and quench surface conditions. Rewetting Experiments like the one in Fig. 6 (Multimedia view) sug-
by definition occurs when a gas-liquid-solid contact line is gest that in quench cooling, immediate rewetting is bound to
formed. Sections III C 1 and III C 2 present and analyze occur on surface temperature below Ttls because it is found that
top views of the jet impingement zone on a highly super- no vapor film exists for Ti = 300 ◦ C. As discussed in Sec. I,
heated steel surface, allowing the interpretation of new inter- a wide range of Trw -values is found, depending on the phys-
face topologies that have not been observed before. The film ical process at hand and on material conditions.21–27 Surface
boiling regime was observed prior to rewetting for initial plate roughness height and wettability have a large effect on the
temperatures of 450 ◦ C–900 ◦ C, even for a high jet subcooling rewetting temperature.28 Quench experiments to show that a
of 80 ◦ C. Rewetting occurred for Ti = 300 ◦ C and a jet sub- vapor film can be formed before rewetting will be discussed
cooling of 30–80 ◦ C without any vapor blanket formation. In in Sec. III C 2; see also Sec. III A.
all cases, a vigorous production of vapor bubbles was observed Just 0.05 ms after the jet has impinged on the surface (stage
within the wetted zone; this is clearly nucleate boiling. In the I in Fig. 6) (Multimedia view), an inner circular wetted area
following, the case of the lowest Ti (300 ◦ C) is considered first exists surrounding by a white narrow band of the rewetting
followed by that of the highest Ti (900 ◦ C). front, as explained in Sec. III A. At the outside edge of the
wetted region, the water is deflected from the surface in a thin
1. Nucleate boiling at low plate temperature (300 ◦ C) sheet (spattering) which eventually breaks up due to surface
Figure 6 (Multimedia view) shows some typical results tension. After 2.8 ms (stage II), the photograph reveals the
of a test for Ti = 300 ◦ C and Tj = 20 ◦ C. The side views are existence of nucleate boiling; many vapor bubbles are growing
at first and collapsing when coming in contact with cool jet
liquid in a later stage.
Bubbles of various sizes are observed, and through the
bigger bubbles, the dry plate area of the bubble foot can be
seen. It is well known that the bubble shape acts as a lens
that a view from the top enables the observation of the dry
plate area of the bubble foot in nucleate boiling.47 It is even
possible to discern in Fig. 6 (Multimedia view) the individual
LED light bulbs via strong reflections near the contact line
of a bubble; the enlargements of Fig. 7 (Multimedia view)
show these light bulb reflections very clearly. The record-
ings of Fig. 7 (Multimedia view) are for Ti = 450 ◦ C and
Tj = 50 ◦ C, just 0.53 s after the jet impingement. The light
bulbs are arranged in a circle with the jet in the center, which
causes the reflections to occur all along the circumference
of the circular contact line of a bubble near the jet center.
It is obvious that the curvature of the vapor-liquid interface
changes strongly near the contact line, which causes strong
light reflections there as a consequence of the Huygens prin-
ciple. Identical reflections have been found and examined in
our laboratory under conditions which facilitated identification
of the contact line.47 The reflections of the LED lights there-
FIG. 6. Quenching for Ti = 300 ◦ C and Tj = 20 ◦ C. Left: stage I, the jet fore prove the existence of contact lines and boiling bubbles.
touches the surface at 300 ◦ C without film boiling; right: stage II, just 2.8 ms
Initially, bubbles are generated at the plate such that conden-
after jet impingement. The side views are artist impressions; the top view
measurements (Multimedia view: https://doi.org/10.1063/1.5054870.1; frame sation occurs at the top of the bubble that is in contact with
rate = 20 kfps; total time of the movie is actually 10.1 ms). fresh approaching water at 20 ◦ C. The continual condensation
122102-9 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

FIG. 7. Typical life cycle of a vapor bubble during nucleate boiling in quench cooling: growth and subsequent collapse by condensation in the subcooled jet,
for Ti = 450 ◦ C and Tj = 50 ◦ C, at t = 0.53 s after the jet impingement. Note the reflections of the individual LED lights near the contact line (Multimedia view:
https://doi.org/10.1063/1.5054870.2; frame rate = 20 kfps; total time of the movie is actually 3.15 ms).

transports heat from the plate to the liquid in between bubbles, 2. Film boiling and bubble formation at high
and because of that a bubble boundary layer with increasing plate temperature (900 ◦ C)
mean temperature develops. As a result, the mean bubble size
increases over time, which is observed in the top view on the Heat transfer phenomena occurring during the early times
RHS of Fig. 6 (Multimedia view). Boiling bubbles are found of the jet quenching process for Ti = 900 ◦ C and Tj = 20 ◦ C
to collapse by condensation after a sudden initial growth, as shown in Fig. 4 are revealed by the corresponding top views
shown in Fig. 7 (Multimedia view), and bubble lifetime is in the shown in Fig. 8. In order to facilitate the understanding of
range 0.3–2.25 ms. This growth and collapse cycle occurs as a interface topologies, Fig. 8 gives both the “Side view” (upper
rule. part) and the corresponding “Top view” (bottom part). The
Summarizing, boiling bubbles observed from above can rewetting process is elucidated by the consecutive stages I to
be recognized from a dark interior and reflections at the con- IV. The sketches “a” and “d” provide an interpretation of the
tact line and the observations of Fig. 6 (Multimedia view) phenomena observed in photographs “b” and “c,” taken from
prove that nucleate boiling occurs during quench cooling for the movie.
Ti = 300 ◦ C and Tj = 20 ◦ C. The occurrence of boiling Directly after touchdown of the jet (t = 0), part of the
bubbles is critical for heat transfer models in jet quenching water becomes vapor and a continuous vapor film is formed.
because the existence of bubbles do alter the liquid film flow This continuous vapor layer remains uninterrupted for more
and convective heat transfer as compared to cooling without than 0.20 s. After t = 0.12 s (stage I-a), the surface tem-
boiling. perature drops only 9 ◦ C (Ts = 887 ◦ C) due to the thermal
The occurrence of boiling is not found by other insulation of the vapor film; the surface temperature has been
experimental studies that could not realize nor utilize top derived from thermocouple readings in the way explained in
views.13,20,19 Karwa and Stephan20 reported wetted sur- Sec. II E. The liquid-vapor interface of the film reflects light
face at 650 ◦ C without boiling activity, for a jet at Tj irregularly and creates the milky appearance in the side-views
= 25 ◦ C and surface heat flux above 6 MW/m2 . Lee et al.19 (stage I-b). This vapor film was observed in all tests for initial
reported the absence of bubbly activity within the wetted surface temperature of 450 ◦ C to 900 ◦ C. It is remarkable that
zone for an initial test plate temperature of 900 ◦ C cooled water jet quenching experiments16–19 reported the absence of
by the round water jet at 20 ◦ C. The authors argued20,19 vapor film blanketing at an initial plate temperature exceeding
that the jet velocity and subcooling can suppress the bub- 700 ◦ C, even at an initial temperature of 1000 ◦ C.17
bly activity in the wetted region. In the measurements per- Note that the interface topologies observed in the top view
formed here with direct visual inspection, this is not the from Fig. 8, stage I-c, do not occur at the low temperature in
case. Fig. 6 (Multimedia view): 300 ◦ C. The existence of whitish
122102-10 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

FIG. 8. Images from top and side views of jet impingement zone during quench test (Ti = 900 ◦ C; Tj = 20 ◦ C; Vj = 1 m/s). Stages: (I) degassing on top of film
boiling, (II) onset of rewetting, (III) rewetting front, and (IV) nucleate boiling.

bubbles, without a dark interior, is revealed in Figs. 8(c) and A comparison between the behavior of a “gas” bubble
8(d) at stage I. These whitish bubbles are found to keep on and a vapor bubble can be made with Figs. 9 (Multimedia
moving quite hectically, without reducing in size, while some- view) and 10, for the same experiment with Ti = 600 ◦ C
times agglomerating without coalescing. These bubbles are and Tj = 50 ◦ C. Figure 9 (Multimedia view) shows succes-
visible until they move out of the observable impingement sive photographs of many “gas” bubbles constantly moving
zone. These bubbles therefore never condensed despite the for 5.8 ms, never condensing, while some are agglomerat-
fact that they are moving in highly subcooled liquid that is ing. These packed bubbles change the position relative to
continually refreshed from the top. This, and arguments given one another but never disappear individually. The zoomed-
below, leads us to believe that these are bubbles filled with in images of two “gas” bubbles, at t = 4.6 and 10 ms, suggest
inert gases, so are effervescence bubbles, although there is no that they are made up of a cluster of smaller “gas” bubbles.
direct proof of that. To verify this assumption, we have var- Apparently coalescence is prohibited by some physical mech-
ied the gas content of the impinging water jet in a subsequent anisms, like Marangoni convection or long liquid film drainage
study in order to examine its effect on mean size of the white times.
bubbles. Indeed we found such an effect, which supports the At 0.12 s after jet impingement (see Fig. 10), the surface
above supposition of effervescence bubbles rather than boiling becomes completely wetted without the presence of “gas” bub-
bubbles. For the sake of conciseness details of this, a follow- bles, but with a multitude of vapor bubbles connected to the
up study will be given in a subsequent manuscript.48 Within surface, increasing in size and subsequently collapsing, with
wetted regions, no “gas” bubbles have been observed. lifetimes less than 0.6 ms. These vapor bubbles possess a dark
122102-11 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

FIG. 9. Top view images for Ti = 600 ◦ C and Tj = 50 ◦ C, just 6.6 ms after jet impingement. The first figure depicts the erratic motion (red lines) of two “gas”
bubbles (red bullet) until they get out of the visible zone based on the top view images that follow underneath. The bubbles are put inside red dashed circles.
The magnified images at t = 4.6 and 10 ms show that the two “gas” bubbles are essentially a cluster of smaller “gas” bubbles, at these times (Multimedia view:
https://doi.org/10.1063/1.5054870.3; frame rate = 5 kfps; total time of the movie is actually 212 ms).

interior and reflections at the contact line.47 The bubbles are bubbles that are pressed onto it. The vertical motion is con-
recognized as vapor bubbles in the same way, as in Sec. III C 1. verted into a sideway motion, which is consistent with the
The rewetting period is further considered in Secs. III D erratic motion of the white bubbles in all kinds of directions.
and III E. Another possible mechanism to prohibit coalescence is the
Whatever the inert gas fraction of the whitish bubbles, agency that also prohibits coalescence of the cluster of bubbles
they appear to be living on top of a vapor film without coa- in Fig. 9 (Multimedia view), discussed above.
lescing with it. This in itself is quite remarkable since the As far as the authors know, the present study is the first
highly subcooled water is pouring down at impinging veloc- one that reports “gas” bubbles occurring on top of a vapor film
ity of 1 m/s onto bubbles and film. The non-coalescence in quench cooling. We do not know how much water vapor
can be explained from the fact that the water-vapor inter- is contained in these “gas” bubbles, but a major part of the
face is mobile and adjusting under the approach of a bubble. content is expected to consist of inert gases that are known to
Thus, this interface acts as a kind of skating surface for “gas” be solved in water at normal conditions.
122102-12 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

FIG. 10. Top view images for Ti = 600 ◦ C and Tj = 50 ◦ C: Vapor bubbles occur in the wetted zone with growth and condensation into the subcooled water.
Note the reflections of the individual LED lights near the contact line. The corresponding movie is linked in the caption of Fig. 9 (Multimedia view).

D. Rewetting mechanisms The wetted area increases in time as the rewetting front
advances outwards [Fig. 8(c), stage III] until the impingement
A continuous film boiling covering the visible impinging zone is completely wetted at t = 0.38 s [Fig. 8(c), stage IV].
zone is observed during the first 0.21 s after the jet impinge- Many vapor bubbles, growing and condensing, are observed
ment (Ti = 900 ◦ C; Tj = 20 ◦ C; Vj = 1 m/s). Then, a small within the wetted region (stage IV-c), where the mean sur-
darker area appears at the center, surrounded by a white nar- face temperature is Ts = 703 ◦ C and surface heat flux is about
row band, indicating that solid-liquid contact (rewetting) took 6 MW/m2 . No “gas” bubbles were observed within this wet-
place. This white band is clearly visible in the top view image ted region. The topology of rewetting is further examined in
in Fig. 8(c), stage II, and demarcates the rewetting front, as Sec. III E.
explained in Sec. III A and shown in Fig. 4. Within this wetted Figure 11 shows the outer surface morphology of the con-
region, nucleate boiling occurs and outside of it, the surface tinuous chromium oxide film grown on the test plate heated
remains dry and covered by a vapor film with many “gas” at 900 ◦ C in air from scanning electron microscope (SEM)
bubbles still existing on top of it. recordings. The test plate material was austenitic stainless
The onset of rewetting occurs at a surface temperature of steel (SS) type 304 with high chromium content (18%-20%)
Ts = 827 ◦ C, which is well above the water critical point, Tcr , that is traditionally known for its resistance against high tem-
of 374 ◦ C. For all experiments from Ti = 450 ◦ C to 900 ◦ C perature oxidation on surface. This resistance is ensured by
and jet temperature Tj = 20 ◦ C to 70 ◦ C, rewetting occurs at the formation of protective chromium oxide (Cr2 O3 ) film.35
surface temperatures above Tcr . This result is in line with our Figure 11(a) shows a global view of the oxidation surface mor-
previous study14 and several jet quenching experiments16–20,29 phology with the presence of a large element of spongy texture
that also reported rewetting at plate temperatures above Tcr . with height about 30 µm. This element may be the result of the
122102-13 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

FIG. 11. Outer surface of the chromium oxide film grown on the test plate during the heating process at 900 ◦ C: (a) Global view of oxidation surface morphology
with the presence of a large element of spongy texture, (b) detailed view of surface chromium oxide crystals that constitute the oxide film on the plate surface,
and (c) “sand rose.”

combustion of residual elements on surface during the heat- Ti = 450 ◦ C–750 ◦ C, oxide layers are observed to be thin-
ing of the plate. Many other types of large elements were also ner. The effectiveness of the oxide layer as a thermal insulator
observed on the heated surface, such as the “sand rose” shown is easily examined as follows. The measured rewetting heat
in Fig. 11(c). The chromium oxide film is constituted of little, flux, qrw , and rewetting temperature, Trw , enable an estimate
regular Cr2 O3 crystal grains, as shown in Fig. 11(b). Those of an effective heat transfer coefficient, h. To determine the
crystals are formed when 304 SS is heated in air to elevated Biot number, the oxide layer thickness is taken as the charac-
temperatures.35 teristic length, L, ranging from 1 to 3 µm, and h = qrw /(Trw
The surface morphology shown in Fig. 11 has cross sec- − Tj ). For Trw = 600 ◦ C, Tj = 20 ◦ C, and qrw = 0.4 MW/m2 ,
tions such as those in Fig. 12. A thin layer of chromium the Biot number ranges from 2.2 × 10−4 to 6.7 × 10−4 with
oxide covers the surface and has a thickness of about 1–2 µm the temperature drop across the oxide layer from 0.13 ◦ C to
[Fig. 12(a)], with peaks reaching heights of 4 µm or more. 0.39 ◦ C, respectively. For Trw = 738 ◦ C, Tj = 20 ◦ C, and
Figure 12(b) shows a cross section of a porous element with a qrw = 0,96 MW/m2 , the Biot number ranges from 4.3 × 10−4
height of about 12 µm. Measurements revealed a total height to 13 × 10−4 with the temperature drop across the oxide layer
roughness, Rt, of 5 µm. The expected vapor film thickness from 0.31 ◦ C to 0.93 ◦ C, respectively. For Trw = 827 ◦ C,
ranges from 3 to 5 µm for a round jet of 10 mm at 20 ◦ C Tj = 20 ◦ C, and qrw = 2.9 MW/m2 , the Biot number ranges from
impinging on a 1000 ◦ C stainless steel plate.17 The surface 17 × 10−4 to 35 × 10−4 with the temperature drop across the
asperities may therefore easily penetrate the vapor film and oxide layer from 0.94 ◦ C to 2.81 ◦ C, respectively. Therefore,
interact with the flowing liquid and, consequently, cool down since Bi  0.1 and the temperature drop across oxide layer is
more rapidly than the remaining solid surface since they act very small, the quenching surface acts as a lumped surface and
as a kind of micro-fin on top of the solid material under- the oxide layer is clearly a negligible thermal barrier. Although
neath. The hydrophilic nature of the solid facilitates wetting it is clearly demonstrated that the drop in temperature through
and film penetration if cooling of the asperities is sufficient. the oxide thickness is insignificant, an increase in this drop is
The cavities created by Cr2 O3 crystals and the porous-spongy noticed with the rise of the Trw .
texture of the large particles, with a high wettability and Figure 13 shows a sequence of events of jet quenching
a high heat transfer area, facilitate rewetting, even at high with some deposit material on the plate for Ti = 900 ◦ C and
mean surface temperatures. This wetting may initiate full wet- Tj = 20 ◦ C; the bonding of the material to the plate is by van der
ting of a part of the substrate at a scale visible from the Waals forces. After a brief period of film boiling (t = 0.02 s),
outside. the rewetting starts (t = 0.23 s) not only in the center but also at
Since the thermal conductivity of oxides is much lower the side (red arrow), exactly on porous dirt deposits. Contrary
than that of metals, it was speculated in the literature8 that to what has been observed in Fig. 4, where the wetted region
rewetting could occur on an oxide surface with a tempera- grows radially outward, there is no rewetting front progressing
ture sufficiently below Ttls , while the surface of the actual from the center. Instead there is “local” rewetting at several
pure metal underneath would be at a high temperature above patches where deposits occur.
the critical point. The conductivity of chromium oxide [3.1 Rewetting starts at higher points on the surface. This has
W/(m K)] is about 10 times lower than that of the test plate. been further investigated by depositing a set of lines of car-
The chromium oxide layer thickness was estimated to range bon powder on the plate before the quench test (see Fig. 14).
from 1 to 3 µm for Ti = 900 ◦ C, as shown in Fig. 12. For Figure 14 shows that in this case, rewetting starts on the

FIG. 12. Test plate heated at 900 ◦ C in


air: (a) cross section of the plate surface
overlaid by the chromium oxide layer
about 1–2 µm with peaks and valleys
and (b) cross section of the plate surface
overlaid by the chromium oxide layer
with the presence of a porous element
12 µm high.
122102-14 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

FIG. 13. Sequence of events of jet quenching with


depositions occurring on the plate; Ti = 900 ◦ C and
Tj = 20 ◦ C.

powder, and Figs. 15(a) (Multimedia view) and 15(b) (Multi- likely because of the variation in powder sizes applied in the
media view) show that rewetting starts on spot corrosion where above experiments.
surface is rougher and more porous. This of course does not The conclusion of the above findings is that rewetting
imply a rigorous proof of similarity with asperities occurring starts at plate asperities that either result from the oxida-
in normal plate surface roughness, but this similarity is highly tion process or from dirt and other irregularities. There is

FIG. 14. Snapshots of the onset of rewetting first on lines of carbon powder (graphite) where surface is more porous for Ti = 450 ◦ C and Tj = 20 ◦ C.
122102-15 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

FIG. 15. Snapshots of the onset of rewetting first on spot corrosion where surface is rougher for Ti = 900 ◦ C and Tj = 20 ◦ C. (a) Multimedia view:
https://doi.org/10.1063/1.5054870.4; frame rate = 10 kfps; total time of the movie is actually 122 ms. (b) Multimedia view: https://doi.org/10.1063/1.5054870.5;
frame rate = 10 kfps; total time of the movie is actually 10.5 ms.

confirmation of these findings in the literature. Based on exper- control of the cooling process. Efficient cooling control with
iments, Bui and Dhir 49 and Sinha et al.50 found that increased elevated cooling rates requires accurate prediction of the onset
roughness or oxidation or deposition on the plate surface of rewetting in terms of temperature and time.3–5,8,17,18,30–34,50
increases the rewetting temperature in quench cooling. They The following analysis leads to a set of correlations that can
explained this by the increase in surface wettability and num- be used in practice.
ber of nucleation sites due to roughness, porosity, and pitting A typical cooling curve with corresponding flow patterns,
of the oxide layer. Also, the spongy texture of oxides would boiling regimes, and rewetting temperature is shown in Fig. 16;
increase the heat transfer area. Filipovic et al.8 suggested that Ti = 751 ◦ C, ∆Tsub = 80 ◦ C, and Vj = 1 m/s. Images A through E
asperities may protrude through the thin vapor layer inter- show the successive stages of flow boiling regimes with names
acting with liquid, inducing a change from film boiling. Our that denote the most striking feature: film boiling, transition,
experiments confirm these explanations. maximum heat flux, nucleate boiling, and single-phase flow.
Film boiling still occurs on the surface 0.02 s after quench-
E. Effect of initial surface temperature ing has started (image A), and the surface temperature, Ts , of
and subcooling on rewetting
751 ◦ C remains unaltered due to the low heat transfer across
Rewetting phenomena are encountered in many manu- the vapor layer.
facturing processes involving quenching of steel on a run-out The red dashed circles in Fig. 16 highlight some of many
table, continuous castings, extrusions, forgings, etc. In hot strip gas bubbles present during film boiling. Rewetting occurs at
rolling mills, where cooling rates and final temperature ensure trw = 0.11 s when surface temperature, Ts , is 738 ◦ C where
its mechanical properties, rewetting is the key requirement for two wetted spots (red dashed circles) are seen in image B.

FIG. 16. Heat transfer and boiling regimes in jet impingement quenching for Ti = 751 ◦ C, ∆Tsub = 80 ◦ C, and jet velocity of 1 m/s.
122102-16 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

Note that rewetting did not occur at the time of a sharp rise in respectively [see Figs. 17(a) and 17(c)]. For ∆Tsub = 80 ◦ C
heat flux but later, at a time when the rewetting heat flux, qrw , with Ti = 750 ◦ C and 900 ◦ C, trw = 0.11 s and 0.21 s and Trw
is 1.1 MW/m2 . Image C shows the wetted area increasing with = 738 ◦ C and 827 ◦ C, respectively [see Figs. 17(b) and 17(d)].
the advancement of the rewetting front outwards (red dashed The images in Fig. 17 clearly show that rewetting may start
line). Intense bubbly activity is observed within the wetted anywhere and even at a place outside of the impingement zone
area, while outside this area, the gas bubbles still remain on [Fig. 17(c)]. The reason for this is explained above: rewet-
top of the vapor film, with Ts = 711 ◦ C and qs = 2.4 MW/m2 . ting can start whenever higher points on the plate surface may
The maximum heat flux, qmax , of 5.0 MW/m2 occurs at penetrate the vapor film after cooling down.
t = 0.30 s and Ts = 574 ◦ C (image D) and marks the beginning An increase in ∆Tsub or Ti causes an increase in Trw . An
of a nucleate boiling regime where the heat flux curve slope increase in ∆Tsub or a reduction in Ti causes a decrease in trw .
becomes negative. The time resolution of Fig. 16 is sufficient These trends are easily understood from the effect of the tran-
to see all the details during rewetting times, and the surface sient driving “force” (temperature differences) in the following
heat flux is found to be at maximum at a time much later than way.
the start of rewetting. The reason for this is that as long as During film boiling, before rewetting occurs, heat transfer
a vapor film covers a substantial part of the surface, so also from the hot surface is by convection and radiation through the
with partial rewetting, the resistance to heat transfer is high. vapor layer to the vapor-liquid interface, if this layer is present.
Only after full rewetting (photograph D), nucleate boiling may The sensible energy of the subcooled water is increased, and
occur anywhere and heat flux may attain a maximum in the the vapor gets slightly superheated. An increase in impinging
same way as it does in a normal pool boiling curve. jet subcooling causes more energy to be delivered to the sub-
A gradual reduction in bubble population is observed until cooled water, via the vapor-liquid interface and at the same
the bubbly activity ceases (Image E) at t = 3.06 s, for Ts time less water to be used for vapor generation. This explains
= 246 ◦ C and qs = 2.1 MW/m2 . Henceforth, the regime is pre- that with increasing jet subcooling, the vapor layer thickness
dominantly single-phase flow. Even though the surface temper- decreases. Similarly, an increase in initial plate temperature
ature of 246 ◦ C is considerably elevated, the high subcooling of increases the vapor generation and the thickness of the vapor
80 ◦ C seems to be the main reason for ceasing the bubbly activ- layer. With decreasing vapor layer thickness, more surface
ity. This suggests that the high subcooling of 80 ◦ C would asperities make contact with the vapor-liquid interface and,
be sufficient to suppress the bubbly activity on nucleation hence, rewetting is promoted. Therefore, an increase in jet sub-
sites. Future work is required for a better understanding of cooling or initial plate temperature causes an increase in the
this phenomenon in jet impingement quenching. rewetting temperature and increase in the rewetting (delay)
time. Quenching experiments conducted by Sinha et al.50
1. Trends explained
produced similar trends.
Figure 17 shows the effects of the initial plate tempera- A raise by a factor of three in jet velocity (1–3 m/s) gave
ture and the jet subcooling, ∆Tsub , on rewetting temperature, only a small change in rewetting temperature (Fig. 18). A simi-
Trw , and rewetting delay time, trw . For ∆Tsub = 50 ◦ C with Ti lar effect of jet velocity on Trw also was also found by Agrawal
= 600 ◦ C and 750 ◦ C, delay time and rewetting temperature et al.18 for Trw ranging from 790 ◦ C to 798 ◦ C and for jet veloc-
were trw = 0.08 s and 0.28 s and Trw = 593 ◦ C and 656 ◦ C, ities ranging from 2 to 9.6 m/s. The jet subcooled was 78 ◦ C,

FIG. 17. Top view images of rewetting for 600 ◦ C ≤ Ti ≤ 900 ◦ C, 50 ◦ C ≤ ∆Tsub ≤ 80 ◦ C, and Vj = 1 m/s.
122102-17 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

to our data the following coefficients have been found:


a1 = 3.00 ± 0.19,
a2 = 1.437 ± 0.016,
a3 = 0.0918 ± 0.046,
and
C1 = 22 563 ± 500,
all with a 95% band of confidence. The correlation can be
used for Ti ≥ 450 ◦ C. The parameters R-square (r2 ) and F-
statistic (F) indicate the quality of the fit. R-square indicates
how successful the fit can explain the variability in observed
data. F is a measure for the appropriateness of the number of
fitting parameters. A high value for F indicates a good choice
of number of parameters to describe the observed data.51 The
FIG. 18. Effect of Ti , subcooling, and jet velocity on rewetting temperature. values found for r2 and F of the above fit are 0.986 and 5409,
respectively, indicating a good fit. It stands to reason that the
inaccuracy of the parameter a3 is high because the number of
and their stainless steel plate was initially at 800 ◦ C. A factor velocities examined is low.
of 5 in velocity resulted in a small increase of 2% in Trw . Also Figure 19 shows a comparison between experimental
Ishigai et al.31 and Ochi et al.29 concluded that jet velocity values and predicted ones of Trw . The predicted values
has a weak effect on Trw . The jet velocity trend affects the have a standard deviation, σ, of ±7.6 ◦ C (95% confidence
thickness of the vapor layer via the hydrodynamic pressure in level).
the impingement zone. This is apparently a minor effect in the The capability of Eq. (1) to predict the rewetting temper-
film boiling regime. ature for test conditions different from those used in present
The experimental results for Trw shown in Fig. 18 and the study is now examined with data from the following studies.
snapshots in Fig. 17 prove that Trw strongly depends on Ti prior Filipovic et al.8 used a copper specimen at 850 ◦ C exposed
to quenching, but that in the range of Ti = 450 ◦ C to 600 ◦ C, to a rectangular water jet (2 ≤ Vj ≤ 4 m/s; 45 ◦ C ≤ ∆Tsub
subcooling has a weak effect on Trw . Also Kokado et al.30 and ≤ 75 ◦ C); Kokado et al.30 used a round impinging jet with a
Filipovic et al.8 found a strong influence of subcooling on Trw . diameter of 10 mm on a 900 ◦ C stainless steel plate; Agrawal
et al.18 used a small jet with a diameter of 2 mm impinging
2. Correlation for rewetting temperature at velocity 2–9.6 m/s on a 800 ◦ C stainless steel plate; Wang
et al.4 used a hot stainless steel plate (200–900 ◦ C) cooled by
Figure 18 shows the experimental Trw values plotted as a round jet (57 ◦ C ≤ ∆Tsub ≤ 87 ◦ C; 2 ≤ Vj ≤ 8 m/s) of 4 mm
function of initial plate temperature, jet subcooling (30 ◦ C–80 diameter. The rewetting temperatures calculated with (1) are
◦ C), and jet velocity (1 and 3 m/s). For T = 450 ◦ C, the symbols
i compared to the values reported in these studies in Fig. 20. The
overlap because rewetting always occurred at Ti , independent predicted values agree well with the experimental results, even
of subcooling and jet velocity. This suggests that Trw = Ti when for plate material and test conditions far out of range of those
Ti is around 450 ◦ C. Up to Ti = 600 ◦ C, Trw is slightly affected
by subcooling and not affected by Vj . For plate temperatures
in the range 600–900 ◦ C, the influence of subcooling on Trw
is remarkable. For the high subcooling of 80 ◦ C, Trw occurs
close to Ti . As Ti increases, Trw increases.
Kokado et al.30 correlated the Trw linearly with Tj for
68 C ≤ Tj ≤ 92 ◦ C, subcooling below 32 ◦ C, and a jet velocity

of 2 m/s. Recently, Takrouri et al.5 and Nobari et al.32 also


proposed correlations for Trw as a function of jet subcooling.
To the best of our knowledge, no correlation for the rewet-
ting temperature of a water jet quenching on a steel surface that
takes the influences of initial surface temperature, subcool-
ing, and impingement jet velocity simultaneously into account
is available in the literature. For this reason, the following
empirical correlation is here proposed:
! a1
C1 Ti − C2
Ti − Trw = a2 a3 . (1)
∆Tsub Vj C2 − Tsat
FIG. 19. Comparison of our experimental Trw data with predictions with
The coefficient C2 can be fixed to 450 ◦ C because data showed C
 T −C  a1
Ti − Trw = a21 a3 C i−Tsat
2 .
that Trw = Ti when Ti is around 450 ◦ C. By least squares fitting, ∆Tsub Vj 2
122102-18 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

FIG. 20. Comparison of literature Trw data with predictions with Ti − Trw
C
 T −C  a1
= a21 a3 C i−Tsat
2 .
∆Tsub Vj 2

FIG. 22. Comparison of experimental values of trw and predictions with trw
used to develop Eq. (1). For jet quenching on stainless steel,   
Ti K2 −K3 Vj 
= K1 exp  b1
 .
the predicted values agree to within ±5%, while for copper  ∆Tsub 
plates, the agreement is to within ±10%.
this reason, the following empirical correlation is proposed:
3. Correlation for rewetting delay
  
Ti K2 − K3 Vj 
The rewetting delay time, trw , is the time interval between trw = K1 exp 
b1
 . (2)
impingement of the jet onto the hot surface and the onset of  ∆Tsub 
rewetting. Figure 21 shows the effect of initial surface temper-
By least mean squares fitting, the following coefficients for
ature, jet velocity, and subcooling on trw . A strong effect of Ti
Eq. (2) were found:
and jet subcooling on rewetting delay time is found, while the
effect of Vj is minor. These results are in agreement with those K1 = 1.53 × 10−3 ± 0.21 × 10−3 ,
of Piggott et al.33
K2 = 36.9 × 10−3 ± 1.8 × 10−3 ,
Existing correlations to predict the rewetting delay time,
trw , in jet quenching are not applicable to the test conditions K3 = 1.92 × 10−3 ± 0.2 × 10−3 ,
of the present study with delay times as short as trw = 0.02 s.
and
Mozumder et al.34 proposed that empirical correlations for trw
b1 = 0.421 ± 0.016,
valid only for rewetting delay times satisfying trw > 30 s and
applicable only outside of the impingent zone (r/dj > 0.5). For all within 95% of confidence. The values found for r2 and F
were 0.998 and 11 430, respectively. Figure 22 shows a com-
parison between the experimental and predicted values of Trw .
The differences had a standard deviation, σ, of 0.012 s (95%
confidence).

IV. CONCLUSIONS
The characterization of the rewetting and boiling phe-
nomenon inside water jet quenching of a high tempera-
ture steel plate was achieved with the aid of a novel high
speed imaging technique. Main findings are summarized as
follows.
Rather surprisingly, non-coalescing gas bubbles have been
observed on top of the vapor film that is formed underneath
the liquid directly after the start of impingement if the plate
temperature is sufficiently high (quantified below).
For the first time, the rewetting process within the jet
impingement zone was directly observed. Deposition of solid
particles on the hot plate, as well as an increased roughness
FIG. 21. Effect of Ti , subcooling, and jet velocity on rewetting delay time, of the plate, induces rewetting. Rewetting occurs on top of
trw . asperities that act as microfins on the hot surface.
122102-19 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

Stable rewetting is found to occur if the plate tempera- and


ture is below the thermodynamic limit of water superheating. Vs r
!
For initial surface temperatures below 300 ◦ C and for a total = 1.33 − 0.0936 , 2.86 ≤ r/dj ≤ 14. (A3)
Vj dj
height roughness of 5 µm, rewetting occurs without a vapor
film covering the plate. For initial surface temperature above These expressions were developed for Reynolds numbers
450 ◦ C, such a vapor film occurs even for a subcooling as high in the range 17 000 < Re < 47 000 and for 2.1 < dj < 9.3 mm.
as 80 ◦ C. However, a jet subcooling of 80 ◦ C is sufficient to The liquid layer thickness (hj ) is well approximated by
suppress this bubbly activity on nucleation sites if the surface hj
! ! !
1 1 1
plate temperature is 250 ◦ C or less. = . (A4)
Surface asperities may easily penetrate the vapor film and dj 5.3 r/dj Vs /Vj
interact with the flowing liquid. By cooling down more rapidly
than the remaining solid surface, they act as a kind of micro-fin 1 D. Hall, F. Incropera, and R. Viskanta, “Jet impingement boiling from a
on top of the solid material underneath. circular free-surface jet during quenching: Part 1—Single-phase jet,” J. Heat
After rewetting of the total surface, intense vapor bub- Transfer 123(5), 901–910 (2001).
2 N. Karwa, L. Schmidt, and P. Stephan, “Hydrodynamics of quenching
ble activity is observed. Vapor bubbles of various sizes are
with impinging free-surface jet,” Int. J. Heat Mass Transfer 55, 3677–3685
observed, and through the bigger bubbles, the dry plate area
(2012).
of the bubble foot can be seen. The reflections of the LED 3 N. Karwa, T. G. Roisman, P. Stephan, and C. Tropea, “A hydrodynamic

lights therefore prove the existence of contact lines and boiling model for subcooled liquid jet impingement at the Leidenfrost condition,”
bubbles. Int. J. Therm. Sci. 50, 993–1000 (2011).
4 B. Wang, X. Guo, Q. Xie, Z. Wang, and G. Wang, “Heat transfer character-
The existence of non-coalescing and whitish bubbles, istics during jet impingement on a high-temperature plate surface,” Appl.
without a dark interior and mentioned above, is found to keep Therm. Eng. 100, 902–910 (2016).
5 K. Takrouri, J. Luxat, and M. Hamed, “Experimental investigation of quench
on moving quite hectically, without reducing in size, while
sometimes agglomerating without coalescing. This and other and re-wetting temperatures of hot horizontal tubes well above the limiting
temperature for solid–liquid contact,” Nucl. Eng. Des. 311, 167–183 (2017).
arguments lead us to believe that these are bubbles filled with 6 J. H. Lienhard, N. Shamsundar, and P. O. Biney, “Spinodal lines and
inert gases, so are effervescence bubbles, although there is no equations of state: A review,” Nucl. Eng. Des. 95, 297–314 (1986).
7 V. Carey, Liquid–Vapor Phase-Change Phenomena, 2nd ed. (Taylor &
direct proof of that.
Temperature and time of rewetting are strongly affected Francis, New York, 2007).
8 J. Filipovic, F. P. Incropera, and R. Viskanta, “Rewetting temperature and
by the initial surface temperature and jet subcooling and less velocity in a quenching experiment,” Exp. Heat Transfer 8(4), 257–270
by the jet velocity. Empirical correlations for these parameters (1995).
9 M. Islam, M. Monde, P. Woodfield, and Y. Mitsutake, “Jet impingement
and dependencies have been presented for plates with a total
height roughness of 5 µm. quenching phenomena for hot surfaces well above the limiting temperature
for solid-liquid contact,” Int. J. Heat Mass Transfer 51, 1226–1237 (2008).
10 P. L. Woodfield, M. Monde, and A. K. Mozumder, “Observations of high
ACKNOWLEDGMENTS temperature impinging-jet boiling phenomena,” Int. J. Heat Mass Transfer
The authors gratefully acknowledge the support of this 48(10), 2032–2041 (2005).
11 M. Ilyas, M. Ahmad, C. P. Hale, S. P. Walker, and G. F. Hewitt, “Rewet-
work by the Brazilian National Council CAPES, Grant No. ting processes during top/bottom re-flooding of heated vertical surfaces,” in
0439-10-0, and are thankful to Paul Bloemen for his help with ASME/JSME 2011 8th Thermal Engineering Joint Conference, Honolulu,
the design and fabrication of the borescope system. Hawaii, 2011.
12 H. Ohtake and Y. Koizumi, “Study on propagative collapse of a vapor film

in film boiling (mechanism of vapor-film collapse at wall temperature above


APPENDIX: MAIN FEATURES OF THE VELOCITY the thermodynamic limit of liquid superheat),” Int. J. Heat Mass Transfer
FIELD NEAR THE STAGNATION POINT 47(8-9), 1965–1977 (2004).
13 M. N. Hasan, M. Monde, and Y. Mitsutake, “Homogeneous nucleation boil-
OF A CIRCULAR JET ing during jet impingement quench of hot surfaces above thermodynamic
limiting temperature,” Int. J. Heat Mass Transfer 54, 2837–2843 (2011).
This appendix summarizes the equations used to construct 14 H. Leocadio, J. Passos, and A. Silva, “Heat transfer behavior of a high
Fig. 2. temperature steel plate cooled by a subcooled impinging circular water jet,”
The radial velocity gradient in the y-direction vanishes30 in 7th ECI International Conference on Boiling Heat Transfer, Florianopolis,
near y/dj = 0.5. Radial velocities in the stagnation region 2009.
15 W. Timm, K. Weinzierl, and A. Leipertz, “Heat transfer in subcooled jet
(0 ≤ r/dj ≤ 0.5) were found to be well represented by the impingement boiling at high wall temperatures,” Int. J. Heat Mass Transfer
correlation " !# 46, 1385–1393 (2003).
Vr y r 16 F. Xu and M. Gadala, “Heat transfer behavior in the impingement zone
= 1.83 − 3.66 . (A1)
Vj dj dj under circular water jet,” Int. J. Heat Mass Transfer 49, 3785–3799 (2006).
17 Z. Liu and J. Wang, “Study on film boiling heat transfer for water jet imping-
For r/dj < 2.5, the maximum radial velocity did not occur ing on high temperature flat plate,” Int. J. Heat Mass Transfer 44, 2475–2481
at free surface but in the layer very near the plate surface (y/dj ≈ (2001).
0.02).29 The maximum radial velocity exceeds the impinging 18 C. Agrawal, R. Kumar, and A. Gupta, “Rewetting and maximum surface

jet velocity (Vj ), in some cases by about 30%. The normalized heat flux during quenching of hot surface by round water jet impingement,”
Int. J. Heat Mass Transfer 55, 4772–4782 (2012).
free-surface velocity, Vs /Vj , was correlated according to 19 S. G. Lee, M. Kaviany, C. Kim, and L. Lee, “Quasi-steady front in quench
! !2 subcooled-jet impingement boiling: Experiment and analysis,” Int. J. Heat
Vs r r
= 0.303 + 0.625 − 0.125 , 0.5 ≤ r/dj ≤ 2.86 Mass Transfer 113, 622–634 (2017).
20 N. Karwa and P. Stephan, “Experimental investigation of free-surface jet
Vj dj dj
impingement quenching process,” Int. J. Heat Mass Transfer 64, 1118–1126
(A2) (2013).
122102-20 Leocadio, van der Geld, and Passos Phys. Fluids 30, 122102 (2018)

21 K. Pandey and S. Basu, “How boiling happens in nanofuel droplets,” 36 F. P. Incropera, D. DeWitt, T. L. Bergman, and A. S. Lavine, Fundamentals

Phys. Fluids 30, 107103 (2018). of Heat and Mass Transfer, 7th ed. (John Wiley & Sons, 2011).
22 P. E. Lapenna, “Characterization of pseudo-boiling in a transcritical nitrogen 37 J. Stevens and B. W. Webb, “Measurements of flow structure in the radial
jet,” Phys. Fluids 30, 077106 (2018). layer of impinging free-surface liquid jets,” Int. J. Heat Mass Transfer
23 M. Yazdani, T. Radcliff, M. Soteriou, and A. A. Alahyari, “A high-fidelity 36(15), 3751–3758 (1993).
approach towards simulation of pool boiling,” Phys. Fluids 28, 012111 38 J. Stevens and B. W. Webb, “Measurements of flow structure in the stagna-

(2016). tion zone of impinging free-surface liquid jets,” Int. J. Heat Mass Transfer
24 P. Kangude, D. Bhatt, and A. Srivastava, “Experiments on the effects of 36(17), 4283–4286 (1993).
nanoparticles on subcooled nucleate pool boiling,” Phys. Fluids 30, 057105 39 J. Stevens and B. W. Webb, “Measurements of the free surface flow structure

(2018). under an impinging, free liquid jet,” J. Heat Transfer 114(1), 79–84 (1992).
25 V. Pandey, G. Biswas, and A. Dalal, “Effect of superheat and electric field 40 J. Taylor, An Introduction to Error Analysis—The Study of Uncertainties in

on saturated film boiling,” Phys. Fluids 28, 052102 (2016). Physical Measurements, 2nd ed. (University of Colorado, 1996).
26 V. Pandey, G. Biswas, and A. Dalal, “Saturated film boiling at various gravity 41 A. Ahmed and M. Hamed, “Modeling of transition boiling under an

levels under the influence of electrohydrodynamic forces,” Phys. Fluids 29, impinging water jet,” Int. J. Heat Mass Transfer 91, 1273–1282 (2015).
032104 (2017). 42 J. Kang, T. K. Kim, G. C. Lee, and H. S. Park, “Minimum heat flux and
27 K. N. Premnath, F. Hajabdollahi, and S. W. J. Welch, “Surfactant effects on minimum film-boiling temperature on a completely wettable surface: Effect
interfacial flow and thermal transport processes during phase change in film of the Bond number,” Int. J. Heat Mass Transfer 120, 399–410 (2018).
boiling,” Phys. Fluids 30, 042108 (2018). 43 N. Bhatt, R. Raj, P. Varshney, and A. R. Pati, “Enhancement of heat transfer
28 H. Kim, B. Truong, J. Buongiorno, and L. W. Hu, “On the effect of rate of high mass flux spray cooling by ethanol-water and ethanol-tween20-
surface roughness height, wettability, and nanoporosity on Leidenfrost water solution at very high initial surface temperature,” Int. J. Heat Mass
phenomena,” Appl. Phys. Lett. 98, 083121-1–083121-3 (2011). Transfer 110, 330–347 (2017).
29 T. Ochi, S. NakanishI, and M. I. S. Kaji, “Cooling of a hot plate with an 44 S. V. Ravikumar, J. M. Jha, and A. M. Tiara, “Experimental investigation

impinging circular water jet,” in Multi-phase Flow and Heat Transfer III of air-atomized spray with aqueous polymer additive for high heat flux
Part A: Fundamentals (Elsevier Science Ltd, Amsterdam, 1984). applications,” Int. J. Heat Mass Transfer 72, 362–377 (2014).
30 J. Kokado, N. Hatta, and J. Takuda, “An analysis of film boiling phe- 45 D. Trujillo and H. Busby, Practical Inverse Analysis in Engineering (CRC

nomena of subcooled water spreading radially on a hot steel plate,” Arch. Press, Boca Raton, 1997).
Eisenhüttenwes. 55(3), 113–118 (1984). 46 N. Hatta, J. Kokado, and K. Hanasaki, “Numerical analysis of cooling
31 S. Ishigai, S. Nakanish, and T. Ochi, “Boiling heat transfer for a plane water characteristics for water bar,” Trans. Iron Steel Inst. Jpn. 23, 555–564
jet impinging on a hot surface,” in Proceedings of 6th International Heat (1983).
Transfer Conference, Toronto, Canada, 1978. 47 C. H. M. Baltis and C. W. M. van der Geld, “Heat transfer mechanisms of a
32 A. H. Nobari, V. Prodanovic, and M. Militzer, “Heat transfer of a stationary vapour bubble growing at a wall in saturated upward flow,” J. Fluid Mech.
steel plate during water jet impingement cooling,” Int. J. Heat Mass Transfer 771, 264–302 (2015).
101, 1138–1150 (2016). 48 K. T. Kempen, H. Leocadio, G. J. M. Priems, C. W. M. Geld, B. van Esch,
33 B. Piggott, E. White, and R. Duffey, “Wetting delay due to film and transition and M. Bsibsi, “Effect of air concentration of water on quench cooling of
boiling on hot surfaces,” Nucl. Eng. Des. 36, 169–181 (1976). steel plates” (unpublished).
34 A. K. Mozumder, M. Monde, and P. L. Woodfield, “Delay of wetting prop- 49 T. D. Bui and V. K. Dhir, “Transition boiling heat transfer on a vertical

agation during jet impingement quenching for a high temperature surface,” surface,” J. Heat Transfer 107, 756–763 (1985).
Int. J. Heat Mass Transfer 48, 5395–5407 (2005). 50 J. Sinha, L. Hochreiter, and F. Cheung, “Effects of surface roughness, oxida-
35 A. C. S. Sabioni, R. P. B. Ramos, V. Ji, and F. Jomard, “About the role of tion level, and liquid subcooling on the minimum film boiling temperature,”
chromium and oxygen ion diffusion on the growth mechanism of oxidation Exp. Heat Transfer 16(1), 45–60 (2003).
films of the AISI 304 austenitic stainless steel,” Oxid. Met. 78, 211–220 51 J. M. Juran and F. M. Gryna, Juran’s Quality Control Handbook (McGraw-

(2012). Hill, New York, 1988).

You might also like