Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics

12-15 June, 2017, Iguazu Falls, Brazil

DEGASSING, BOILING AND REWETTING IN FREE SURFACE JET


QUENCHING
H. Leocadio*a,b, C. W. M. van der Geld*b, J. C. Passosc
a) R&D Center, Usiminas Steel, Ipatinga, Brazil, hormando.leocadio@usiminas.com; b) Department of Chemical
Engineering and Chemistry, Eindhoven University of Technology, Eindhoven, Netherlands, C.W.M.v.d.Geld@tue.nl;
c) Department of Mechanical Engineering, Federal University of St Catarina, Florianopolis, Brazil.

Abstract. The present study applies high-speed imaging (20 kfps) all through the water jet impinging vertically on a hot
steel plate during quench cooling. Top views of boiling phenomena within the impingement zone have been obtained
for the first time, in this way. The quenching of 300-900°C hot steel plates by a water jet of 20-70°C has been studied.
The inverse heat conduction method gave the heat flux and the temperature at the plate surface from temperatures
measured with thermocouples inserted in a steel plate. For the first time, gas bubbles, from a degassing process, have
been observed on top of a thin vapor film. Roughness induces rewetting on the surface at local top plate temperatures
well above the critical point of water. For a surface at 300 °C, rewetting occurs without the occurrence of film boiling,
but for surfaces above 450°C and even in high subcooling this is not the case. Suppression of bubbly activity on a
surface at 246 °C was observed for high jet subcooling. Rewetting temperature and rewetting delay time are strongly
affected by the initial surface temperature and by the subcooling.

Keywords: Degassing, Boiling, Rewetting, Quenching of steel, Jet impingement

1. INTRODUCTION

Jet impingement boiling has high cooling potential and it is used as an effective mean of cooling in many industrial
applications involving rapid cooling, temperature control in emergency core cooling of nuclear power plants, and power
dissipation in the micro-electro-mechanical devices. Metal industries widely employ it for accurate temperature control
and ensure the mechanical and metallurgical properties, where the surface temperature (~ 900°C), heat flux
(above 6 MW/m²) and cooling rate (above 100°C/s) are typically very large and acceptable cooling times are relatively
short. After decades of wide utilization, the complete heat transfer process remains not fully understood due to the
existence of simultaneously different boiling regimes, which change over time, on impingement surface which depend
on the surface temperature, geometry, subcooling, and flow conditions of jet (Takrouri et al., 2017; Hasan et al., 2011).
Since, each boiling regime promotes large differences in heat removal capacities and the lack of understanding of
thermal behavior of has limited the development of effective cooling schemes (Wang et al., 2016).
Rewetting (solid-liquid contact) is a critical factor for the design and accurate control of quench-based
manufacturing process and issues pertinent to nuclear reactor safety because the rapid cooling and maximum heat flux
are obtained when it is established, what has conducted a number of studies in the last decades in quenching of steel by
water jet impingement (Ishigai et al., 1978; Kokado et al., 1984; Islam et al., 2008; Karwa & Stephan 2013; Paul et al.,
2016). Rewetting temperature is the maximum surface temperature at which solid-liquid contact can occur and has
different physical meanings of Leidenfrost temperature which is the maximum temperature at which an isolated droplet
floating on a vapor cushion eventually collapses and touches the surface (Filipovic et al., 1995; Carbajo, 1985).
However, in several studies (Agrawal et al., 2012; Filipovic et al., 1995; Liu & Wang, 2001; Takrouri et al., 2017;
Leocadio et al., 2009) with water jet impingement quenching, rewetting has been measured at surface temperatures well
above 700°C that greatly exceed the water critical point of 374°C, which is the limit condition for water remains liquid
(Carey, 2007). Hall et al. (2001) advised those high rewetting temperatures above 900°C reported by Ishigai et al.
(1978) and Ochi et al (1984) and 710°C by Filipovic et al. (1995) should be considered physically impossible.
Thermodynamic limit of liquid superheat (T tls) is the maximum upper limit of superheating of the liquid without
undergoes instantaneous vaporization (Carey, 2007). For Monde’s group (Hasan et al., 2011; Islam et al., 2008;
Woodfield et al., 2005; Mozumder et al., 2005; Monde, 2008) and other authors (Ilyas et al., 2011) the solid-liquid
interface could never exceed thermodynamic limit of water superheat around 306°C. According to them, neither high
speed visual observation nor temperature measurement could find out concrete explanation about what happens in the
early stages of water jet impingement quenching, when the surface temperature remains well above the critical
temperature of water. Another unexpected finding reported in literature (Xu & Gadala, 2006; Liu & Wang, 2001; Ochi
et al., 1984; Ishigai et al., 1978) was the rewetting taking place without being preceded by any period of time of film
boiling, in despite of surface temperature as high as 1000°C and jet velocity and high subcooling can suppress the
bubbly activity. Wang et al (2016) investigated the heat transfer characteristics in water jet (13- 43°C) impingement on
a hot steel plate (200-900°C) and concluded the rewetting temperature highly depend on the initial surface temperature,
slightly the water jet temperature and almost unaffected by the jet velocity. Water jet temperature had little effect on
their heat flux curves. Contrary to conclusions of Wang et al (2016), Takrouri et al (2017) concluded initial surface
9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

temperature has no effect on rewetting temperature from their investigation of quench of (380‒780°C) horizontal tubes
by jet impingement. In others studies (Karwa & Stephan, 2013; Karwa et al., 2012), although the surface temperature in
the wetted region is high enough for nucleate boiling and heat flux above 5 MW/m², no bubbly activity were identified.
Karwa et al. (2012) pointed out, at this stage, it is not known whether or not any bubbles exist in the wetted region, and
if they exist, do they alter the liquid film flow, what is a critical aspect for the development of models for jet
impingement quenching.
According to the survey above mentioned, ambiguities and contradictions do exist among researchers and the
boiling phenomena in the jet impingement quenching remain unknown without concrete evidences that explain
rewetting phenomenon on surface temperature beyond critical point of water. The present work conducted an
experimental study of water jet impingement quenching on a high temperature steel plate, recording, for the first time,
through the water jet, yielding top view high-speed images (up to 20 kfps) of impingement zone. The effects of
subcooling, jet velocity and initial test temperature on the heat transfer regimes were analyzed by means of the transient
inverse heat conduction method what predict the heat flux and surface temperature on the impinging surface from
temperatures measured with thermocouples inserted in test plate. A stainless steel plate preheated at 300°C to 900°C
was quenched by a round water jet of 8 mm, ranging 20°C to70°C, with velocity of 1 and 3 m/s.

2. EXPERIMENTAL APPARATUS AND PROCEDURES

2.1. Experimental set up with borescope

Figure 1. (a) Schematic of the test rig; (b) Schematic of cylindrical water jet impinging on test plate with
thermocouples indicated as TC1, TC2 and TC3 and with (y, r) coordinate system definition.

The main components of the experimental apparatus is outlined in Fig. 1-a. The coolant fluid used was the ordinary
city water stored in an open head tank. The desired water temperature is maintained constant in the head tank by a
thermostat controller. The centrifugal pump sends water from bottom tank (reservoir) to the head tank. The overflow
tube had the purpose of keeping constant the water level in the head tank, assuring a constant pressure in water box and,
consequently, a stable circular water jet at nozzle exit which was located centrally at 40 mm from the test surface. The
nozzle has a measured inner diameter of 9.7 mm. The borescope had its base connected to the objective lens of the high-
speed video camera #1 while the viewer was inserted into the water box. The borescope viewer was positioned centrally
above the hole of nozzle exit in manner not disturb the water jet stability and to allow recording high-speed imaging
through the water jet, yielding top views of all stages of quenching phenomena in the jet impingement zone on hot
plate. High-speed camera #1 captured images at range of 5 to 20 kfps. The second high-speed camera captured images
by side of the jet at 300 fps. A series of LED lights surrounding the nozzle exit supply enough lighting to record the
inner events at impingement zone. The water flow rate is set by needle valve. The cross section of the circular
impinging jet of 8 mm on plate test of 50 x 50 x 10 mm³ is presented shown in Fig. 1-b at same dimensional scale.
The hydrodynamic parameters of the impinging jet at center of plate surface (r = 0), such as, jet velocity at the
nozzle exit (Vn), impinging jet velocity (Vj), impinging jet diameter (dj) and the saturation temperature (T sat) at the
9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

impingement zone are shown in Table 1 and they were calculated using the equations of continuity and Bernoulli. The
stagnation pressure, Pj, represents the active pressure in the central point (r = 0) of top plate surface where occurs the
collision of the jet at Vj. For practical reasons, saturation temperature at stagnation point was considered 100°C in the
present study. The jet centrally impinges on the plate (r = 0) at velocity of Vj. Three holes, 1.1 mm in diameter, were
drilled with a depth of 9 mm and checked by calipers. The thermocouples TC1, TC2 and TC3 are located in line at
radial positions of 0, 9 and 18 mm, respectively, from the center of plate at depth of 1 mm below the top surface.
Grounded thermocouples Type-K, 1 mm in sheath diameter made in same material of the test plate to reduce different
thermal behavior inside the plate, was used. High temperature thermal paste with conductivity of 70 W/(m.K) was
inserted into the hole to insure good thermal contact between the thermocouple and steel plate. The test plate material
chosen was the type 304 stainless steel (SS) which produces protective chromium oxide (chromia) layer against high
temperature oxidation, avoiding the formation of peeling of oxide scale, and eliminate the heat generation that occurs in
carbon steel and distracts in the temperature reading. On the sides of the plate without water impingement, insulation
with low conductivity of 0.11 W/(m.K) at 800°C and 25 mm thick was used.

Table 1. Hydrodynamic parameters at stagnation point (r = 0).


Qn dn Vn dj Vj Pj Tsat
(ℓ/min) (mm) (m/s) (mm) (m/s) (kPa) (°C)
3 9.7 0.7 8 1 102 100
9 9.7 2.0 8 3 106 101

The thermal properties of the 304 SS (Leocadio et al., 2009) are shown in Table 2. The test plate was heated until
the test temperature by a manual electrical oven (3 kW) at test quenching position. The data acquisition system
simultaneously triggers the automatic valve, the high-speed camera, LED lights and acquires and stores the data. The
plate temperature history during cooling process was measured by three thermocouples inserted into the test plate at rate
of 50 Hz. The test plate was heated 50°C beyond the initial test temperature, at test position. The quenching tests have
carried out at initial test temperatures of 300, 450, 600, 750, and 900°C which are commonly used metal industrial. The
water was heated up by electrical heater and recirculated through head tank, water box and reservoir until reach the
desire test temperature. The water jet temperatures were 20, 50 and 70°C at 1 and 3 m/s. Before each experiment, the
impinging surface was sanded with 320 grit sandpaper and cleaned with acetone. The measured arithmetic mean
roughness (Ra) at impingement zone surface, after quenching test, was found ranging 0.11 to 0.18 µm. For the new
plate surface Ra = 0.10 µm.

Table 2. Thermophysical properties of the 304 SS.

T (ºC) cp (J/kg.K) ρ (kg/m3) k (W/m.K)


27 447 7900 15.2
127 515 7859 16.6
327 557 7774 19.8
527 582 7685 22.6
727 611 7582 25.4
927 640 7521 28.0

Results of an uncertainty analysis of the primary measurements (k, r, t, y, T m, Tj, Vn) and propagation of these
uncertainties into the calculated results are presented in Table 3. All uncertainties were calculated to be within a
confidence interval of 95%. Those uncertainties were analyzed according to criteria suggested in reference (Taylor,
1996) and account for errors in measurement, calibration, machining, and measuring devices. Combining the
uncertainties of thermocouple and data acquisition system, the total uncertainty of temperature measurement system
was ± 0.30% within 95% limits. Although uncertainty in the thermocouple depth position is small, the effect of these
measurements on calculated heat fluxes and surface temperatures is significant, reaching 60% in the estimation of
maximum heat flux on quenching surface for measurement position difference of 0.6 mm. The thickness of weld
closure covering the tip of thermocouple, where the hot joint is grounded, usually varies from 0.1 to 1.0 mm depending
on the manufacturer. To minimize the uncertainty of the measurement position and ensured a higher accuracy in the
estimation of heat flux, the thermocouples were opened and its weld closure thickness has been measured at the end of
tests. The calculated relative uncertainty for heat flux (q) was ± 5 %.
9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

Table 3. Relative uncertainties in calculated and measured quantities.

Quantity Relative or Total Uncertainty


r δr = ± 0.05 mm
t δt = ± 0.01 s
k δk / k = ± 6 %
y δy = ± 0.05 mm Measured
Tm δTm / Tm= ± 0.3%
Tj δTj = ± 2°C
Vn δVn / Vn= ± 3 %
q δq / q = ± 5 % Calculated

2.2. Inverse heat conduction analysis

The internal temperature history of the test plate registered by thermocouples was used to predict the unknown heat
flux and temperatures on the quenching surface. To estimate the heat flux and temperature distribution along the
quenching surface, a successfully applied (Leocadio et al., 2009; Mohapatra et al., 2013; Ravikumar et al., 2014; Lee et
al., 2004; Ahmed & Hamed, 2015) commercial inverse heat conduction software, INTEMP, was used varying the
thermal properties of material. The detailed methodologies of modeling and computations were described by Trujillo
and Busby (1997). INTEMP uses the dynamic programming method to solve the nonlinear inverse heat conduction
problem with finite element methodology and it uses the Crank–Nicolson formulation to solve the heat conduction
model, because of its accuracy and least square method, together with a regularization term used to minimize the error
in estimated fluxes. Hence, it solves for the unknown flux histories at any number of specified locations. The input data
consist of a geometrical field (nodes and elements), temperature-dependent thermal properties and the measured time–
temperature histories.

Figure 2. 2D axisymmetric finite element model divided into three unknown heat flux monitoring zones on top plate
surface and location of thermocouples

A 2D axisymmetric finite element model was used in the present study for the numerical analysis. The model had
25 mm radius, 10 mm thickness with 551 and 1000 quadratic elements of 0.5 x 0.5 and 0.5 x 0.2 mm², respectively,
with 4-nodes per element as showed in Fig. 2. The 551 small elements filled the first 4 mm from the top surface with
finer elements in the vertical direction in order to enhance the numerical analysis, because of fast time response needed
close to quenching surface. For boundary conditions, the faces of the plate without jet impingement were considered
adiabatic. The impingement surface was divided into three unknown heat flux zones: r1 = 0 to 5 mm (Zone 1); r2 = 5 to
14 mm (Zone 2); r3 = 14 to 25 mm (Zone 3). The measured temperature histories are given as input at the nodes
corresponding to the position of the thermocouple TC1, TC2 and TC3 located at 1 mm from top surface at radial
position of r = 0, 9 and 18 mm, respectively. Temperature-dependent thermal properties of 304 SS, shown in Table 2,
were used in the present numerical model to avoid the large errors found in predicted heat flux when constant thermal
properties are used (Karwa et al., 2012). The validation of this model was verified against ANSYS-commercial
engineering simulation software of finite element analysis. The mean errors found between those calculated surface
temperatures by ANSYS and predicted surface temperature by INTEMP-IHC model were 0.01%, 0.01% and 0.01%,
respectively.
9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

3. RESULTS AND DISCUSSION

3.1. Gas and vapor bubbles

Of particular interest are the early stages (0 ≤ t ≤ 0.38 s; t = 0 at jet touch-down) of quenching of a hot flat plate.
Fig. 3 shows a typical example for a plate at initial surface temperature, Ti, of 896°C by, jet velocity of 1 m/s and jet
subcooling of 80K. Naturally, film boiling occurs immediately, but solid-liquid contact (rewetting) and nucleate boiling
also occur, leading to stages I, II, III and IV, as defined in the images a, b, c, and d. The sketches "a" and "d" provide an
interpretation of the phenomena observed in photographs "b" and "c", based on the movie, which warrants easy
interpretation. In the top view c (stage I) bubbles are discerned as white spots. Due to the facts that these bubbles never
condense, while some coalesce, and that they all keep on moving hectically until they move out of the observed
impingement zone, it is concluded that these bubbles consist of inert gas rather than vapor. Further evidence for this
interpretation and more details of this phenomenon are presented in another study (Kempen et al., 2017) where it is
proven that these gas bubbles originate from degassing of the impinging liquid. As far as the authors know, this is the
first time that air bubbles were visualized during quenching.

Figure 3. Images from top and side views of impingement zone during quenching by a subcooled jet at 80K
and 1 m/s of a plate initially at 896°C. Stages: (I) degassing on top of film boiling, (II) start of rewetting, (III)
rewetting progresses, and (IV) nucleate boiling.

Stage I-a clearly shows that part of the impinging water becomes vapor and that film boiling starts immediately. The
side image of Stage I-b shows the vapor layer as a white disk, which is confirmed by the top view in Stage I-c. After
0.12 s (Stage I-a) the surface temperature has reduced by only 9°C due to the thermal insulation of film boiling. The
film boiling was observed during a period of 0.21 s even for high subcooling of the impinging jet. This finding
9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

contradicts some recent studies (Liu & Wang, 2001; Agrawal et al., 2012; Karwa et al., 2012; Xu & Gadala, 2006) and
former studies (Ishigai et al., 1978; Ochi et al., 1984), which reported that high subcooling and a high jet velocity could
suppress film boiling. This makes clear that direct observation through the impinging jet is quite elucidating.
After 0.21 s, rewetting takes place at the center of the impingement zone; the surface temperature in the center is
then 827°C, which is quite above the water critical point of 374°C; see Stage II-c. This high rewetting temperature is in
line with most studies mentioned in the above but not with two (Ilyas et al., 2011; Hasan et al., 2011) that suggested that
the rewetting interface temperature cannot exceed the thermodynamic limit of liquid superheat of 306°C.
Film and nucleate boiling coexist within the impinging zone during Stage III (starting at t = 0.27 s; Ts = 800°C). The
wetted area increases with time due to the advancement of rewetting front outwards. Many bubbles and slugs of vapor
growing and collapsing are observed, while outside of wetted zone gas bubbles still remain on top of the film boiling.
In Stage IV (starting at t = 0.38 s; Ts = 703°C) the impingement zone is completely wetted and many vapor bubbles
growing and condensing with maximum diameters around 0.5 mm are observed, while the heat flux is about 6 MW/m².
This boiling activity was not found in some previous jet quenching studies (Karwa & Stephan, 2013; Karwa et al.,
2012; Liu & Wang, 2001; Ishigai et al., 1978) with high temperature SS plate cooled by a high subcooled jet. Instead,
these studies presumed suppression of boiling activity because of the high jet velocity and the high subcooling of the
jet, despite the heat flux of 6 MW/m². Out of the impingement zone in Stage IV–b the film boiling still exists and the
wetted area continues to grow.
Figure 4 shows some typical effects of the initial surface temperature (Ti) and jet subcooling (ΔTsub) on rewetting
temperature (Trw) and rewetting delay time (trw). For a subcooling of 50K with Ti = 600°C (Fig. 4-a) and 750°C (Fig. 4-
c), trw = 0.08 s and 0.28 s and Trw = 593°C and 656°C, respectively. For ΔTsub = 80K with Ti = 750°C (Fig. 4-b) and
896°C (Fig. 4-d), trw = 0.11 s and 0.21 s and Trw = 738°C and 827°C, respectively. An increase in subcooling and/or
initial surface temperature causes a raise in rewetting temperature. An increase in subcooling and/or reduction in initial
surface temperature decrease the rewetting delay. Therefore, both rewetting delay and rewetting temperature are
affected by initial surface temperature and jet subcooling. The first of these results contradicts Takrouri et al (2017) who
reported that the initial surface temperature would have no effect on the rewetting temperature; a similar jet subcooling
effect was found by Takrouri et al. (2017). Our results also agree with those of with Wang et al (2016) who investigated
the heat transfer characteristics in water jet (13°C-43°C) impingement boiling on a hot steel plate (200°C-900°C) and
concluded the rewetting temperature highly depend on the initial surface temperature and slightly of jet subcooling.

Figure 4. Top views images of rewetting for four initial surface temperatures in the range 600°C - 900°C; jet
subcooling is 50K to 80K and jet velocity is 1 m/s.

For all experiments with an initial surface temperature of 450°C to 900°C and subcooling of 30K to 80K, the
rewetting temperature was above and beyond the water critical point of 374°C; this finding agrees with several quench
studies mentioned above. However, the general expectancy was that once the rewetting temperature was reached,
rewetting would occur in the whole impingement zone immediately. This was not the case. The images in Fig. 4-clearly
show that rewetting may start anywhere and even at place out of the impingement zone (Fig. 4-c). Rewetting may start
9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

whenever an elevated point becomes wet; this presumption implies that the surface condition and roughness in
particular, affects this phenomenon. This will be thoroughly examined in section 3.2.

Figure 5. Quenching for Ti = 300°C and ΔTsub = 80K, (a) jet touches the surface at 300°C without film boiling; (b)
just 2.8 ms after jet collision.

A quenching test for a jet subcooling of 80K at 1 m/s impinging on a 300°C-preheated surface was carried out.
Fig. 5 shows the jet touching down the hot surface (Stage I) without film boiling formation. Just 0.0028 s after jet
impingement (Stage II), the surface is wetted and a high bubble activity is observed. On the contrary, film boiling
formation was observed in all experiments with Ti = 450°C to 900°C. This suggests that at surface temperatures around
the thermodynamic limit of water superheat, Ttls, of 306°C no film boiling occurs and solid-liquid contact is bound to
occur. This was also suggested by several other studies (Islam et al., 2008; Woodfield et al., 2005; Hasan et al., 2011;
Ilyas et al., 2011). The thermodynamic limit of liquid superheat, Ttls, is the upper limit of superheating of the liquid
without yielding instantaneous vaporization (Carey, 2007). If a solid-liquid interfacial temperature equals or exceeds the
Ttls then the liquid is repelled from the hot surface and rewetting cannot occur (Filipovic et al., 1995).
Physical causes of rewetting will be further examined in section 3.2 and the corresponding temperatures and heat
fluxes in section 3.3.

3.2. Roughness for boiling and rewetting

The surface morphology of 900°C plate quenched by a jet, 80 K subcooled, is shown in Fig. 6. Fig. 6-A shows the
protective chromium oxide (chromia) covering the surface and the presence of an impurity with spongy texture and
estimated height of 30 µm, 200 times higher than the measured roughness, Ra. The Ra parameter is the arithmetic mean
of the absolute values of the roughness profile and does not give an indication of the extremes, the peaks and valleys on
the solid surface. Therefore, the total height roughness (Rt) is found to be 20 times as large as the Ra. Others types of
impurities also were found (not shown here). Fig. 6-B shows a detail of chromia crystals overlaying the surface. Fig. 6-
C shows cross-section of the test plate. The chromia thickness is estimated to be in the range 1 - 2 μm. The porous
9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

impurity found has a large dimension that exceeds by far the thickness of the chromia layer. Experiments with initial
plate temperatures less than 900°C yielded less oxidation and a thinner chromia layer. The thermal barrier created by
the chromia layer has been calculated. A difference of 1.8°C between the oxide and steel surfaces was found for
Ti = 900ºC and for an oxide layer of 2 µm. For Ti = 750°C the difference was 0.5°C for an oxide layer of 1 µm. Even
though the heat conductivity of chromia is 10 times less than that of stainless steel 304, both surfaces have about the
same temperature. Quenching tests with different surface conditions were carried out, with results shown in Fig. 7. In
all tests, rewetting started exactly where the asperities and impurities occurred.

Figure 6. Surface morphology of a 900°C-preheated plate quenched by 80K-subcooled jet: A) oxidized surface; B)
chromia crystals; C) cross-sections of the test plate.

Figure 7. Images of the onset of rewetting on surface asperities.


9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

The sketches on the right hand side (RHS) of Fig. 7 outline the surface condition at the start of the quenching
experiment. Fig. 7-a shows a surface at 450°C with straight scribbles of graphite and some corrosion spots (see red
arrows) before being impacted by a subcooled jet at 80K at time t = 0. The white blurs surrounding the edge of the jet
are reflections of the LED lights. Just after 9 ms, the rewetting starts exactly on the corrosion spots and on the scribbles
of graphite (red arrows) where surface is rougher with a porous morphology. Fig. 7-b shows a surface at 600°C with
dust smudges just after jet impingement at time zero; subcooling is 80K. Again, rewetting occurred exactly on the
smudges where roughness is highest. Fig. 7-c shows rewetting occurring on corrosion spots (surrounded by a red dashed
line) for a preheated surface at 900°C; subcooling is 80K. This experiment shows the occurrence of rewetting on
different locations at the same time, as also shown in Fig. 4.
Liu and Wang (2001) studied jet impingement on a stainless steel surface at 1000°C by a circular jet with a diameter
of 10 mm at 2 and 6 m/s. Subcooling was 80K and vapor film thickness was found to be 5 and 3 µm. The impurity
shown in Fig. 6 is about 30 µm high, which exceeds 3 µm. It is easily seen that asperities or impurities can penetrate
through the vapor layer and establish contact with the flowing liquid (Fig. 8), but if asperity temperature is higher than
thermodynamic limit of water superheat, then, the liquid is repelled from asperity surface (Fig. 8-a). The upper part of
such an asperity or impurities would be cooled down to temperatures close to the thermodynamic limit of water
superheat of 306°C, allowing for rewetting (Fig. 8-b). Such asperities or impurities would act as microfins removing
heat from hot surface to the liquid. A higher roughness would yield a higher rewetting temperature and a higher heat
transfer rate, if this interpretation were correct.

Figure 8. Schematic representation of rewetting and boiling induced by asperities.

The boiling and rewetting mechanism explained in Fig. 8 cools the plate surface temperature first locally to the
thermodynamics limit temperature, Ttls, and later full solid-liquid contact takes place. Thus, asperities operate as
microfins on the plate surface and rewetting occurs first on upper part of asperities. This explains the high rewetting
temperatures measured in both in the present study and in the literature. Increasing the initial surface temperature
increases rewetting delay because the time spent for asperity reaching Ttls. Both the delay time and the temperature of
rewetting increase when subcooling and/or initial surface temperature are increased, as shown in Fig. 4. On a really
smooth surface rewetting is expected to occur on a plate with a mean temperature fairly close to T tls. Sinha et al. (2003)
indeed showed by their boiling curves that the rewetting temperature increased with an increase of surface roughness,
however no significant effect on rewetting temperature was observed with an increase of initial surface temperature.
According to Paul et al (2016) roughness and the initial surface temperature affect the rewetting temperature. The heat
fluxes and flow patterns connected to rewetting are examined in the next section.

3.3. Cooling curves

A typical cooling curve with corresponding flow patterns and boiling regimes are shown in Fig. 9; Ti = 751°C,
ΔTsub = 80K and Vj = 1 m/s. Images A through to E show the successive stages of flow boiling regimes with names that
denote the most striking feature: film boiling, transition, maximum heat flux, nucleate boiling, and single-phase contact
flow. Film boiling still occurs on the surface 0.02 s after quenching has started (image A) and the surface temperature,
Ts, of 751°C remains about unaltered due to the low heat transfer across vapor layer. The red dashed circles in Fig. 9
highlight some of many gas bubbles present during film boiling. Rewetting occurs at trw = 0.11 s when surface
temperature, Ts, is 738°C where two wetted spots (red dashed circles) are seen in image B. Note that rewetting did not
occur at the time of a sharp rise in heat flux but later already at a time when the rewetting heat flux (qrw) is 1.1 MW/m².
This is different from the finding reported by Wang et al. (2016). Image C shows film boiling, transition boiling and
nucleate boiling regimes coexisting at the surface while rewetting progresses. The wetted area increases over time with
the advancement of the rewetting front outwards (red dashed line). Intense bubbly activity is observed within the wetted
area while outside this area the gas bubbles still remain on top of the vapor film, with Ts = 711 °C and a surface heat
9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

flux, qs, of 2.4 MW/m². The maximum heat flux, qmax, of 5.0 MW/m² occurs at t = 0.30 s and Ts = 574°C (image D) and
marks the beginning of a nucleate boiling regime where the heat flux curve slope, the time rate of change of the heat
flux, becomes negative. The gradual reduction in bubble population is due to the surface temperature reduction, which
causes a reduction of the number of active nucleation sites. Bubbly activity ceases and a predominantly single-phase
flow regime starts at t = 3.06 s, for Ts = 246°C and qs = 2.1 MW/m² (image E).

900 Film boiling Transition Nucleate Boiling 6


A B qmax D Single
Surface Temperaure,°C

800 Phase 5

Heat Flux, MW/m²


700
Trw 4
600
C 3
500 Tmax E
Surface Temperature
2
400 Heat Flux
1
300

200 trw 0

100 -1
0.0 0.1 Time, s 1.0
Figure 9. Heat transfer and boiling regimes in jet impingement quenching for T i = 751°C, ΔTsub = 80K, and
jet velocity of 1 m/s.

The bubbly activity ceases because a minimum surface superheat must be attained for any cavities on the surface to
become active (Carey, 2007). A higher subcooling corresponds to a higher surface temperature to initiate nucleation.
Kandlikar (2006) studied nucleation during subcooled flow boiling in microchannels and suggested Eq. (1) to estimate
the minimum surface superheat required for nucleation:

𝑇𝑂𝑁𝐵 − 𝑇𝑠𝑎𝑡 = √(8.8𝜎𝑇𝑠𝑎𝑡 𝑞𝑠 )/(𝜌𝑣 ℎ𝑙𝑣 𝑘𝑙 ) (1)

Here, , ρv, kl, hlv, qs are the surface tension coefficient (N/m), vapor mass density (kg/m3), thermal conductivity
(W/mK), heat of vaporization (J/kg) and surface heat flux (W/m²), respectively. Substituting the appropriate properties
in Eq. (1), TONB is found to amount 242°C, which is close to the 246°C found in the present experiment. This
observation shows that the high subcooling of 80K suffices to suppress the bubble activity on nucleation sites.
Future manuscripts will detail other heating and boiling characteristics during quench cooling of hot steel plates.

4. CONCLUSIONS

The characterization of flow patterns and boiling phenomena during the quenching of high temperature steel plate
by a circular subcooled water jet was achieved with the aid of a novel high speed imaging technique. The following
main conclusions are drawn:
- For the first time, gas bubbles from degassing have been observed on top of a vapor film that covers the hot plate.
- Roughness and surface condition make that rewetting occurs at temperatures well above the critical point of water.
Asperities act as microfins on the superheated surface and rewetting occurs on top of those asperities.
- For surface temperatures below 300°C rewetting occurs without a vapor film. Data suggest that Ttls is the
maximum surface temperature at which solid-liquid contact is possible. For plate temperatures at or exceeding 450°C,
jet velocity and high subcooling cannot suppress film boiling to occur prior to rewetting.
9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

- The rewetting temperature and the rewetting delay time are strongly affected by the initial surface temperature and
by the subcooling.
- Suppression of bubbly activity was observed to occur on surface at 246°C. High subcooling and jet velocity cannot
suppress bubbly activity on surface temperatures at and above 300°C.

5. ACKNOWLEDGEMENTS

The authors gratefully acknowledge support of this work by the Brazilian Coordination for the Improvement of
Higher Education Personnel (CAPES), Proc. 405700/2013-0. They also express their gratitude to ing. Paul Bloemen
for his great contribution in the development of the optical measuring technique.

6. REFERENCES

Agrawal, C., R. Kumar, and A. Gupta. "Rewetting and maximum surface heat flux during quenching of hot surface by
round water jet impingement." International Journal of Heat and Mass Transfer, 2012: 4772–4782.
Ahmed, A.B., and M.S. Hamed. "Modeling of transition boiling under an impinging water jet." International Journal of
Heat and Mass Transfer, 2015: 1273–1282.
Carbajo, J. J. "A study on the rewetting temperature." Nuclear Engineering and Design, 1985: 21–52.
Carey, V.P. Liquid–Vapor Phase-Change Phenomena. Second Edition. New York: Taylor & Francis, 2007.
Filipovic, J., F. P. Incropera, and R. Viskanta. "Rewetting temperature and velocity in a quenching experiment."
Experimental Heat Transfer, 1995: 257–270.
Hall, D.E., F.P. Incropera, and R. Viskanta. "Jet Impingement Boiling From a Circular Free-Surface Jet During
Quenching: Part 1 - Single-Phase Jet." ASME Journal of Heat Transfer, 2001: 901-910.
Hasan, M. N., M. Monde, and Y. Mitsutake. "Homogeneous nucleation boiling during jet impingement quench of hot
surfaces above thermodynamic limiting temperature." International Journal of Heat and Mass Transfer, 2011:
2837–2843.
Ilyas, M, M Ahmad, C P Hale, S P Walker, and G F Hewitt. "Rewetting Processes During Top/Bottom Re-Flooding of
Heated Vertical Surfaces." ASME/JSME 2011 8th Thermal Engineering Joint Conference. Honolulu, Hawaii, 2011.
Ishigai, S., S. Nakanish, and T. Ochi. "Boiling heat transfer for a plane water jet impinging on a hot surface."
Proceedings 6th Int. Heat Transfer Conference. Toronto, Canada, 1978. 445-450.
Islam, M.A., M. Monde, P.L. Woodfield, and Y. Mitsutake. "Jet impingement quenching phenomena for hot surfaces
well above the limiting temperature for solid-liquid contact." International Journal of Heat and Mass Transfer,
2008: 1226–1237.
Kandlikar, S G. "Nucleation characteristics and stability considerations during flow boiling in microchannels."
Experimental Thermal and Fluid Science, 2006: 441–447.
Karwa, N., and P. Stephan. "Experimental investigation of free-surface jet impingement quenching process."
International Journal of Heat and Mass Transfer, 2013: 1118–1126.
Karwa, N., L. Schmidt, and P. Stephan. "Hydrodynamics of quenching with impinging free-surface jet." International
Journal of Heat and Mass Transfer, 2012: 3677–3685.
Van Kempen, K T., Leocadio, H, Priems, G J M; van Geld, C W M; van Esch, B; Bsibsi, M. "Effect of air
concentration in the water on quench", submitted, 2017.
Kokado, J., N. Hatta, and J. Takuda. "An Analysis of Film Boiling Phenomena of Subcooled Water Spreading Radially
on a Hot Steel Plate." Archiv für das Eisenhüttenwesen, 1984: 113-118.
Lee, P., H. Choi, and S. Lee. "The Effect of Nozzle Height on Cooling Heat Transfer from a Hot Steel Plate by an
Impinging Liquid Jet." Iron and Steel Institute of Japan International, 2004: 704–709.
Leocadio, H., J.C. Passos, and A.F.C. Silva. "Heat transfer behavior of a high temperature steel plate cooled by a
subcooled impinging circular water jet." 7th ECI International Conference on Boiling Heat Transfer. Florianopolis,
2009.
Liu, Z., and J. Wang. "Study on film boiling heat transfer for water jet impinging on high temperature flat plate."
International Journal of Heat and Mass Transfer, 2001: 2475-2481.
Mohapatra, S. S., S. V. Ravikumar, R. Ranjan, S. K. Pal, and S. B. Singh. "Ultra Fast Cooling and Its Effect on the
Mechanical Properties of Steel." Journal of Heat Transfer, 2013: 1-9.
Monde, M. "Heat Transfer Characteristics during Quench of High Temperature Solid." Journal of Thermal Science and
Technology, 2008: 292-308.
Mozumder, A. K., M. Monde, and P. L. Woodfield. "Delay of wetting propagation during jet impingement quenching
for a high temperature surface." International Journal of Heat and Mass Transfer, 2005: 5395–5407.
Ochi, T., S. NakanishI, and M. Ishigai, S Kaji. "Cooling of a hot plate with an impinging circular water jet." Edited by
T.N. Veziroglu and A.E. Bergles. Multi-Phase Flow and Heat transfer III. Part A: Fundamentals. Amsterdam:
Elsevier, 1984. 671-681.
9th World Conference on Experimental Heat Transfer, Fluid Mechanics and Thermodynamics
12-15 June, 2017, Iguazu Falls, Brazil

Paul, G, PK Das, and I Manna. "Assessment of the process of boiling heat transfer during rewetting of a vertical tube
bottom flooded by alumina nanofluid." International Journal of Heat and Mass Transfer 94 (2016): 390–402.
Ravikumar, S V, J M Jha, and I Sarkar. "Mixed-surfactant additives for enhancement of air-atomized spray cooling of a
hot steel plate." Experimental Thermal and Fluid Science, 2014: 210–220.
Sinha, J., L.E. Hochreiter, and F.B Cheung. "Effects of surface roughness, oxidation level, and liquid subcooling on the
minimum film boiling temperature." Experimental Heat Transfer: A Journal of Thermal Energy Generation,
Transport, Storage, and Conversion, 2003: 45-60.
Takrouri, K, J Luxat, and M Hamed. "Experimental investigation of quench and re-wetting temperatures of hot
horizontal tubes well above the limiting temperature for solid–liquid contact." Nuclear Engineering and Design 311
(2017): 167–183.
Taylor, J.R. An Introduction to Error Analysis – The Study of uncertainties in physical measurements. Second Edition.
University of Colorado, 1996.
Trujillo, D.M., and H.R. Busby. Practical Inverse Analysis in Engineering. Boca Raton: CRC Press, 1997.
Wang, B, X Guo, Q Xie, Z Wang, and G Wang. "Heat transfer characteristics during jet impingement on a high-
temperature plate surface." Applied Thermal Engineering 100 (2016): 902–910.
Woodfield, P L, M Monde, and A K Mozumder. "Observations of high temperature impinging-jet boiling phenomena."
International Journal of Heat and Mass Transfer, 2005: 2032–2041.
Xu, F., and M.S. Gadala. "Heat transfer behavior in the impingement zone under circular water jet." International
Journal of Heat and Mass Transfer, 2006: 3785–3799.

You might also like