You are on page 1of 15

Ironmaking & Steelmaking

Processes, Products and Applications

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/yirs20

Experimental investigation of heat transfer


characteristics during water jet impingement
cooling of a high-temperature steel surface

H. Leocadio & J. C. Passos

To cite this article: H. Leocadio & J. C. Passos (2021): Experimental investigation of heat
transfer characteristics during water jet impingement cooling of a high-temperature steel surface,
Ironmaking & Steelmaking, DOI: 10.1080/03019233.2021.1872467

To link to this article: https://doi.org/10.1080/03019233.2021.1872467

View supplementary material

Published online: 24 Jan 2021.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=yirs20
IRONMAKING & STEELMAKING
https://doi.org/10.1080/03019233.2021.1872467

Experimental investigation of heat transfer characteristics during water jet


impingement cooling of a high-temperature steel surface
a b
H. Leocadio and J. C. Passos
a
Research & Development Division, Usiminas Steel, Ipatinga, Brazil; bDepartment of Mechanical Engineering, Federal University of Santa Catarina,
Florianopolis, Brazil

ABSTRACT ARTICLE HISTORY


The microstructure and mechanical properties of the hot rolled steel depend on the laminar jet Received 15 October 2020
cooling control on the runout table. An experimental and numerical study of the heat transfer Revised 15 December 2020
during water jet impingement cooling of a hot steel plate (600 to 900°C) is presented. The effects Accepted 4 January 2021
of plate temperature upon the heat transfer were analyzed by the inverse heat conduction
KEYWORDS
method that provided the heat fluxes on top surface from measured temperatures into the plate. Hot steel rolling mill; steel
Rewetting (liquid-solid contact) occurs on surface above 880°C. Higher initial temperatures enlarge temperature control; jet
the heat flux, but delay the onset of the rewetting and the advance of rewetting front radius impingement cooling; heat
(Rwet). Heat transfer coefficient (HTC) increases as surface temperature decreases. The size of the transfer; accelerated cooling
non-symmetric effective jet cooling zone depends on strip speed. Correlations to predict HTC and system; rewetting
Rwet have been proposed. Results will contribute for enhancement of fast cooling system in hot temperature; inverse heat
strip mill. conduction method; heat
transfer coefficient; effective
cooling zone ; water jet
cooling

Introduction throw to reach the bottom strip surface, configured to


Jet impingement boiling has high cooling potential and it is match the heat transfer from the upper surface. Nozzle diam-
used as an effective mean of cooling in many industrial appli- eter ranges from 10 to 20 mm.
cations involving rapid cooling, temperature control in emer- In this process, strip temperatures and cooling rates are
gency core cooling of nuclear power plants, and power extremely large and cooling times are very short, which
dissipation in the micro-electro-mechanical systems devices makes it important to obtain accurate heat transfer coeffi-
[1]. Metal industries widely employ water jet impingement cients and classify heat transfer modes, in order to model
as a means of precisely controlling temperature during and design efficient cooling control systems [6,8–10].
cooling process of hot rolling, continuous casting, extrusions, However, the heat transfer in the water jet impingement on
forging and hot strip mill cooling on run-out tables [2–4]. The high-temperature surfaces is complex due to the existence
temperature control at the runout table is one of the most of different boiling regimes (film boiling, rewetting, nucleate
critical processes to ensure required mechanical properties and transition boiling) simultaneously at different streamwise
and grain structure in a steel strip. High-quality steels with locations on a moving surface [3,5–7,9].
less expensive alloys may be obtained with accelerated Rewetting is the process of establishing direct liquid water
cooling process by impinging water jet [2,5–7]. Accelerated contact with the hot surface [4,11–13]. Rewetting tempera-
cooling is generally achieved using sprays, water curtains ture (Trw) is the surface temperature that allows the solid–
(planar jets) and round jets (circular jets). Round jets liquid contact establishment. Rewetting delay (trw) is the
provide greater uniformity in cooling over the length of the time spent to rewetting occurs after the jet has impinged
accelerated cooling zone, thereby compensating for their on the hot surface. The elevated heat fluxes and cooling
lower specific cooling performance relative to a planar jet rates start only after the rewetting has taken place [13–15].
system [7]. For a hot strip mill travelling up to 20 m s−1 on Therefore, an effective cooling control starts at the onset of
runout table, the finishing and the coiling temperatures are rewetting. The accurate knowledge of Trw and trw is essential
around 900°C and 150–750°C, respectively, depending on for the enhancement and development of ultra-fast cooling
steel grade. In order to achieve the desired coiling tempera- technology [13,15,16]. After the rewetting under the jet
ture, the strip is cooled simultaneously on top and bottom impingement zone, a visible dark (wetted) circle appears
surfaces by the several arrays of impinging circular water under the jet where high cooling efficiency occurs [11,13,17].
jets installed along the runout table. In order to minimize Some water jet quenching studies [1,18] argued that
temperature variations in the lateral direction on the top of rewetting can occur only on surface below critical tempera-
the strip, U-Tube top headers with double rows of staggered ture of water of 374°C. However, other works [13,19–21]
arrays of round jets are used (see Figure 1) [5,7]. Bottom reported rewetting surface above 700°C without prior film
headers have a single row of laminar jets with sufficient boiling formation, despite initial test temperature (Ti) as

CONTACT H. Leocadio hormando.leocadio@usiminas.com hljunior@yahoo.com.br Research & Development Division, Usiminas Steel, Ipatinga, 35160-900,
Brazil
Supplemental data for this article can be accessed at https://doi.org/10.1080/03019233.2021.1872467
© 2021 Institute of Materials, Minerals and Mining
2 H. LEOCADIO AND J. C. PASSOS

Figure 1. Hot strip mill cooling system with various arrays of impinging circular water jets installed along the run out table and a sketch a U-Tube top header with
double rows of staggered arrays of jets.

high as 1000°C. Ochi et al. [22] and Ishigai et al. [21] reported sometimes agglomerating without coalescing. They never
that water jet (Tj) below 60°C, impinging on 1000°C-steel condensed despite the fact that they are moving in highly
surface, prevents the vapour film formation prior to rewetting subcooled liquid that is continually refreshed from the top
take place and strongly affects the heat flux. Liu and Wang (details in [11]). A small darker (wetted) area appears at the
[23] observed rewetting occurring on surface at 1000°C centre (stage II-c), surrounded by a white narrow band, indi-
without vapour film formation, for water jet diameter of cating that solid–liquid contact (rewetting) took place and
10 mm (Tj = 20°C; Vj = 2–6 m s−1) impinging on steel speci- surface temperature is 827°C. This white band is clearly
men preheated at 1000°C. Takrouri et al. [16] reported Trw is visible in the top view image and demarcates the rewetting
not affect neither Ti nor jet velocity (Vj), but Trw is strongly front. The wetted area increases in time as the rewetting
dependent on Tj, from jet quenching tests on horizontal front advances outwards (stage III-c), Ts = 800°C and vapour
tubes (380–780°C). The influence of surface roughness on bubbles can be visualized within wetted area, until the impin-
rewetting temperature and rewetting delay was recently gement zone is completely wetted at t = 0.38 s (stage IV-c).
reported by Wang et al. [15] and other authors [4,24]. Many vapour bubbles are growing and condensing, within
Recently, in author’s previous work [11] was applied high- the wetted region (stage IV-c), where Ts = 703°C and no
speed imaging (20,000 fps) all through the 10 mm-circular ‘gas’ bubbles were observed within this wetted region.
water jet (20–70°C) impinging vertically on a hot steel plate This study [11] elucidated the rewetting mechanism on
(300⍰900°C) during quench cooling, where rewetting temp- such high-temperature surfaces by examining the surface
erature and delay time were carefully examined and quan- conditions by imagens through the impinging water jet on
tified. Top views of boiling phenomena, within the a steel plate. The limit condition for water remains liquid is
impingement zone, allowed a clear direct observation of the critical temperature (Tcr) of 374°C, at pressure of
the vapour film formation and rewetting phenomenon, fol- 22 MPa. The thermodynamic limit of water superheat
lowed by vigorous boiling activity. Rewetting mechanism, (without undergoes instantaneous vapourization) is Ttls =
which allows rewetting on surfaces above 800°C, was 325°C, at atmospheric pressure [1,25,26]. Thus, if the solid–
reported. Images from top and side views of jet impingement liquid interface exceeds 325°C, the liquid have to be repelled
zone during cooling of a steel plate (Ti = 896°C) by a circular from the hot surface and rewetting cannot occur. Vapour film
jet (Tj = 20°C) are shown in Figure 2 [11]. formation prior rewetting was visually confirmed for all exper-
To facilitate the understanding of interface topologies, iments (Ti = 450–900°C) in [11], except for Ti = 300°C where
Figure 2 gives both images the ‘Side view’ (upper part) and rewetting occurred immediately in the whole impingement
the corresponding ‘Top view’ (bottom part). The rewetting zone, without vapour film formation, flowed by bubbly
process is elucidated by the consecutive stages I to IV. The activity. This suggested that rewetting is bound to occur on
sketches lines ‘a’ and ‘d’ provide an interpretation of the surface temperature below Ttls. Hence, jet quenching tests
phenomena observed in photographs ‘b’ and ‘c’. Just after for different surface conditions was carried out. For all tests
touchdown of the jet (t = 0), part of the water becomes rewetting started first where the surface was more rusty
vapour and a continuous vapour film is formed. This continu- and with highest elevated points on surface, was visually
ous vapour layer remains uninterrupted for more than 0.20 s. confirmed.
After t = 0.12 s (Stage I-a) the surface temperature drops only Since the rewetting temperature at interface solid–liquid
9°C (Ts = 887°C) due to the thermal insulation of vapour film. cannot exceed Ttls, the probable rewetting mechanism on
The liquid–vapour interface of the film reflects light irregu- surface temperatures above Ttls is explained in Figure 3 [26].
larly and creates the milky appearance in the side-views Initially, the steel plate, asperities and oxides have temperature
(Stage I-b). The whitish bubbles (stage I-c & d) are ‘gas’ above Ttls and when the water jet touches the hot surface the
bubbles from the degassing process which keeps on liquid is repelled and a vapour film is formed (Figure 3(a)). Over
moving quite hectically, without reducing in size and time, the upper part of higher asperities is cooled below Ttls
IRONMAKING & STEELMAKING 3

Figure 2. Images from top and side views of jet impingement zone during quench test (Ti = 896°C; Tj = 20°C; Vj = 1 m s−1). Stages: (I) gas bubbles on top of film
boiling, (II) onset of rewetting, (III) rewetting front, and (IV) nucleate boiling [11].

Figure 3. Schematic representation of rewetting and boiling induced by asperities [26].

(Figure 3(b)). Asperities penetrate through the vapour layer the surface temperature. This explanation is in line with Filipo-
and interact with the flowing subcooled water, inducing a vic et al. [4]. Paul et al. [27], Bui and Dhir [28], Wang et al. [15]
change from film to transition boiling regime and remaining and Sinha et al. [24] found that roughness or oxidation or
solid surface > Ttls (Figure 3(b)). Such asperities will act as deposition on the plate surface increase wettability and
microfins removing heat (q) from hot surface to the liquid number of nucleation sites, causing an earlier collapse of
(Figure 3(b)). This rewetting process spreads on vicinity asperi- vapour film, enhanced rate of heat transfer and enlargement
ties, then, more asperities would work as a fin reducing faster of rewetting temperature.
4 H. LEOCADIO AND J. C. PASSOS

The characterization of heat transfer regime within impin- correlations for HTC and advance of rewetted front were pro-
gement zone is critical for the development of jet heat trans- posed. Images show the shape of the effective cooling zone
fer models, since the existence or not of bubbles boiling in (wetted zone) on moving hot strip at run out table.
the liquid film flow can alter the jet cooling efficiency [14].
Authors disagree on the heat transfer regime in impinging
zone during the cooling of the hot strip steel. Some authors Experimental apparatus and procedures
[5,6,9,29,30] suggested that, under impinging jet on a
moving hot strip, the single-phase (without boiling activity) Experimental facilities
regime occurs and should be used in cooling models. Con- An outline of the experimental facilities at R & D Center of Usi-
trary, Timm et al. [31] argued that boiling activity regime minas is shown in Figure 4. The coolant was ordinary water
have to occur due to high heat flux and surface temperature stored in a tank (16), whose equivalent pressure is 70 kPa,
and the vapour bubbles formed on the hot surface condenses creating a water jet flow (13) at the exit (14) of a U-tube
into subcooled liquid. Karwa and Stephan [14] reported (radius of curvature 70 mm), with an inner diameter of
wetted surface at 650°C without bubble boiling activity, for 10 mm. U shaped tube (14) keeps the water a stable flow
Tj = 25°C and the surface heat flux (qs) above 6 MW m−². Lee state, without the presence of air and compatible with ones
et al. [13] also reported the absence of bubbly activity used the cooling system of the steel industry. The overflow
within the wetted zone for a 900°C-plate cooled by 20°C-jet. tube (1) keeps the water level in the tank (16) constant,
They [13,14] argued that Vj and jet subcooling suppress the thus assuring a constant pressure in the U-tube. The steel
bubbly activity in the wetted region. However, boiling activity plate (9) was heated to the test temperature in a 75 kW elec-
was observed on wetted steel surface at 250°C to 800°C, and trical furnace, 800 mm wide, 1000 mm long and 600 mm tall,
ceasing below 250°C, in impingement zone by 20–70°C-jet capable of heating up to 1300°C.
[11]. Furthermore, results published [7,13,19] show discrepan- The test plate material is made of traditional austenitic stain-
cies among heat transfer coefficient (HTC) values. [7] devel- less steel type 304 in order to avoid the heat generation,
oped a mathematical model to predict the thermal caused by phase transformation common in carbon steel,
behaviour of steel strips cooled by array of round jets. They and it resistances against high-temperature oxidation on
[7] considered effective heat removal occurs in wetted circle surface, which is ensured by the formation of protective chro-
(dark circle) under jet impingement equal to 2 Dj (impinging mium oxide (Cr2O3) film [32]. Since Cr2O3 has a very low
jet diameter) and depends on surface temperature (Ts), Vj and thermal conductivity of 3.1 W m−1 K−1, the temperature differ-
Tj. Cox et al. [5] considered a single-phase regime within ence between the surfaces of Cr2O3 and steel, during rewetting
impingent zone equal to 3 Dj and a constant HTC, in their process, was evaluated in previous jet quench study [11] based
strip cooling model. Wiskel et al. [6] quantified a non-circular on Biot number (Bi). Those test plates where heated in air (oxi-
impingent zone 2 Dj long and Dj wide, using infrared camera dizing atmosphere) in similar conditions of the present study.
in a moving surface at 3–8 m s−1. Then, a rectangular dark The temperature gradient across Cr2O3 layer will be insignifi-
zone with a fixed heat flux was considered in their finite cant if Bi < 0.1 [33]. Cross-section of the test plate surface over-
element model for strip cooling. laid by chromium oxide layer was examined. The thin oxide
This study presents the heat transfer behaviour during the layer thickness was estimated ranges from 1 to 3 μm, for Ti =
cooling of a steel plate heated to 600, 750 and 900°C by an 900°C (worst condition). Biot number ranged from 0.002 to
impinging circular water jet. These test temperatures are 0.003 with the temperature drop across the oxide layer from
common found in hot steel mill lines. The effects of Ti on 0.9°C to 2.8°C, respectively. For lower Ti = 450–750°C oxide
qs, HTC, and rewetting were analysed by transient inverse layers were observed to be thinner. Therefore, since Bi <<
heat conduction method that predicts the heat flux and 0.1, the quenching surface acts as a lumped surface and the
temperature on the top surface from temperature histories oxide layer clearly is a negligible thermal barrier. However,
measured in the test plate. A digital camera was used to thick oxide layer, lager than 30 μm, could act as thermal insula-
capture the images during the cooling process. Empirical tion and influence the heat transfer in quenching process [34].

Figure 4. Outline of the experimental jet cooling apparatus.


IRONMAKING & STEELMAKING 5

Thermal properties of the 304 SS for temperatures aligned at distances of 0, 15, 35 and 55 mm, from the
between 27 and 927°C are shown in Table 1. Two kinds of centre of the 150 × 150 mm2 plate, and 0, 15, 35, 55 and
square plates of 150 × 150 × 14 mm³ and 200 × 200 × 75 mm, from the centre of the 200 × 200 mm2. All thermo-
10 mm³ were used. On the sides of the plate without water couples were inserted into by holes drilled from the bottom
impingement, insulation material was used with a thermal of the plate. The tip of the all thermocouples were placed 4
conductivity of 0.170 W m−1 K−1 (at 1090°C) and mm below the top plate surface. Grounded thermocouples
0.078 W m−1 K−1 (at 540°C) and 50 mm thickness. The data K-type, 1.5 mm in sheath diameter made of 304 SS (the
acquisition system, composed of 10 channels, allowed for same as test plate), are used. The actual location of the ther-
an acquisition rate of 20 Hz. mocouple junction is determined by cutting the thermo-
couple after its use in the experiments and by measuring
the distance to the tip of the thermocouple. The heat
Uncertainty analysis
diffusion of the material around the measuring junction is
Results of an uncertainty analysis are presented in Table 2, the same as that of the steel plate and the distance of this
were k, rm, t, ym, Tm, Tj, Vn, Vj and qs are thermal conductivity, junction to the surface is precisely known.
radial position, time, axial coordinate, temperature, water jet High temperature thermal paste with conductivity of
temperature, and water jet velocity at nozzle exit, respect- 70 W m−1 K−1 was inserted into the hole to insure good
ively. All uncertainties are within 95% limits and account for thermal thermocouple-plate contact. The test plate was
errors in measurement, calibration, machining, and measur- heated in furnace to the test temperature and transferred
ing devices. For the temperature measured internally in the to the test position. The cooling process started at test temp-
plate, Tm, sheathed grounded K-Type thermocouples have eratures of 600, 750 and 900°C. The water jet exit was set at a
been used that provided a calibration accuracy of ±1.1°C or height H of 300 mm from the plate surface. The water jet
±0.4% of reading (ASTM E230), whichever is greater, with temperature, Tj, was 22°C, measured using a K-type thermo-
uncertainties within 99% limits. As the test temperatures couple (6) located inside the header (12). A constant air temp-
were 600–900°C, the relative value of ±0.4% produces a erature of 25°C was assumed. A digital camera was used to
greater uncertainty in measured temperatures than ±1.1°C capture the images during the cooling process. Before each
and was therefore selected for the uncertainty analysis. The experiment, the impinging surface was sanded with 320
data acquisition system produces a mean error of ±0.17% grit sandpaper and cleaned with acetone. The measured
within 99% of confidence level. Combining the uncertainties arithmetic mean roughness (Ra) at impingement zone
of thermocouple and data acquisition system, the total uncer- surface, after quenching test, was found ranging 0.15–
tainty of temperature measurement system was ±0.43% 0.26 μm. For the new plate surface Ra = 0.11 μm.
within 99% limits or ±0.30% within 95% limits. The water
flow rate of the jet, Qn, was 6 ℓ min−1, measured with a
flowmeter with a built in measurement error of ±0.25% and Hydrodynamic parameters in water jet
repeatability of ±0.1%. All uncertainties (measured and calcu- impingement
lated) in Table 2 are within 95% limits and are analysed
Some hydrodynamic parameters are required in a heat trans-
according to criteria suggested by Taylor [35]. These criteria
fer analysis, such as: the jet velocity at the nozzle exit (Vn), the
take errors in measurement, calibration, machining and
impinging jet velocity (Vj), the jet diameter just before impin-
measuring devices into account.
gement (Dj) and the saturation temperature of water (Tsat) at
the impingement zone. These parameters are listed in Table 3
Experimental procedure and they were calculated using the equationsof
 continuity

and Bernoulli: Vj = Vn2 + 2gH, Dj = Dn Vn /Vj , and
The plate temperature history during cooling was measured Pj = Patm + (rj .Vj2 )/2, where Dn is the nozzle diameter, H is
with thermocouples(11) inserted into the plate mounted nozzle exit-to-plate surface distance equal to 300 mm, ρj is
the mass density of water, Pj is stagnation pressure, Patm is
Table 1. Thermophysical properties of the 304 SS [33]. atmospheric pressure, and g is the gravitational acceleration.
T (°C) cp (J kg−1 K−1) ρ (kg m−3) k (W m−1 K−1) The jet velocity at nozzle exit, Vn, is calculated from Vn = 4Qn/
27 447 7900 15.2 (πD2n ). The tests were carried out at an altitude above sea level
127 515 7859 16.6 of 234 m, where the atmospheric pressure was Patm ≈
327 557 7774 19.8
527 582 7685 22.6
97.2 kPa. The stagnation pressure, Pj, represents the active
727 611 7582 25.4 pressure on top surface at the centre of the plate (r = 0)
927 640 7521 28.0 where Pj > Patm. The saturation temperature (Tsat) at r = 0
was found to be practically equal to 100°C. The saturation
temperature, Tsat, must be carefully considered in water jet
Table 2. Uncertainties. quenching cooling, because higher Tsat anticipate the
Quantity Uncertainty change from nucleate boiling to the single-phase regime,
rm δrm = ±0.05 mm contributing to an enhancement in the cooling capacity.
t δt = ±0.01 s Impinging water jet temperature (Tj) was 22°C, i.e. subcooling
k δk / k = ±6%
ym δym = ± 0.05 mm
Tm δTm / Tm= ±0.3% Table 3. Hydrodynamic parameters at stagnation point, r = 0.
Tj δTj = ±2°C
Vn δVn / Vn = ±2% Tsat
Vj δVj / Vj = ±3% Qn (ℓ min−1) Dn (mm) Vn (m s−1) Dj (mm) Vj (m s−1) Pj (kPa) (°C)
qs δqs / qs = ±5% 6.0 10 1.3 6.8 2.7 101 99.9
6 H. LEOCADIO AND J. C. PASSOS

of 78°C. Water jet subcooling (ΔTsub) is given by quenching studies [11,39–41] and it allows the thermal prop-
DTsub = Tsat − Tj . erties of material to vary with temperature in each finite
Heat transfer in jet impingement cooling of steel plates is element. This is an important characteristic because it
strongly dependent on how water flows over the surface, avoids errors about 20% due to the assumption of constant
which will be seen to affect the dynamics of the wetting thermal properties in estimating the surface heat flux in jet
front. For this reason, water jet impingement hydrodynamics quench cooling experiments [42]. Temperature-dependent
on an unheated flat surface are briefly described. Based on thermal properties of 304 SS, shown in Table 1, were used
studies of Stevens and Webb [36–38], profiles of the normal- in the present analysis. INTEMP uses the dynamic program-
ized radial velocity, Vr/Vj, across the liquid layer are given for ming method to solve the nonlinear inverse heat conduction
several radial positions, r/Dj, in Figure 5 together with a problem with the finite element method. It uses the Crank –
photograph (Figure 5(a)) of the jet diameter contraction as Nicolson formulation to solve the heat conduction model,
it goes down. together with a regularization term used to minimize the
As the circular free-surface jet issuing from a pipe-type error in estimated fluxes. The temperature boundary con-
nozzle impinges onto the surface, the flow is diverted radially dition at bounding surfaces other than the top of the plate
outwards along the surface and its free liquid jet layer thick- is prescribed. The surface heat fluxes assumed at the top of
ness, hj, is reduced, as shown in photograph (Figure 5(a)) of the plate are subsequently modified by a nonlinear optimiz-
the circular water jet on a cold flat plate surface. The jet diam- ation technique to minimize the error between the measured
eter, Dj, just before impingement at the surface is smaller than temperatures and the predictions. Convergence is reached
jet diameter at nozzle exit, Dn, due to the gravity which once the heat flux reaches the optimum value. The method-
increases the velocity of the water. As the circular free- ologies of modelling and computations were detailed by Tru-
surface jet impinges onto the surface, the flow is spreads jillo and Busby [43]. Two 2D axisymmetric finite element
out radially outward reducing its liquid layer thickness, hj. models were used for the numerical analysis. The first
The pressure is maximum at the stagnation point (r = 0) due model had a radius of 75 mm, 14 mm thickness and 4200
to the dynamic contribution of the impinging jet and quadratic elements of 0.5 × 0.5 mm² with 4-nodes per
decreases monotonically to the ambient value away from it element for the plate 150 × 150 mm², as shown in Figure 6.
at about r/Dj ≈ 1.28 [22]. Conversely, the radial velocity (Vr) The second model had a radius of 100 mm, 10 mm thickness
is zero at the stagnation point (r = 0; y = 0) and increases to and 4000 quadratic elements of 0.5 × 0.5 mm² for the plate
the jet velocity (Vj) with increasing distance along the 200 × 200 mm².
surface. For r/Dj < 2.5, the maximum radial velocity (Vr) did Plate faces without impingement were considered adia-
not occur at free surface but in the layer very near the plate batic, since the radiation and free convective heat transfer
surface (y/Dj ≈ 0.02). The maximum radial velocity exceeds rate are much smaller than that impinging side due to
the impinging jet velocity (Vj), in some cases by about 30%. thermal insulation assembled. The top surface of the square
plate of 150 × 150 mm² was divided into four zones with a
uniform heat flux in each zone: r1 = 0–7 mm (zone 1), r2 = 7–
Inverse heat conduction analysis 25 mm (zone 2), r3 = 25–45 mm (zone 3) and r4 = 45–75 mm
A commercial inverse heat conduction programme, INTEMP, (zone 4). For the plate of 200 × 200 mm² there is an added
was used to predict the heat flux and temperature distri- region, r5 = 75–100 mm (zone 5). The tip of the all thermo-
bution along the quenching surface from temperature his- couples were placed 4 mm below the top plate surface. In
tories measured with thermocouples inserted in the test order to increase the accuracy in the estimation of heat
plate. INTEMP has been applied successfully in several flux, the actual location of the thermocouple juction is

Figure 5. Profiles of the normalized radial velocity, Vr/Vj, in the liquid layer thickness (hj) at various radial positions, r/Dj and (a) photograph of a circular water
impinging jet of Dj = 6.8 mm and Vj = 2.7 m s−1 at 22°C.
IRONMAKING & STEELMAKING 7

Figure 6. 2D axisymmetric finite element model divided into four unknown heat flux monitoring zones on top of the plate surface and location of thermocouples
TC1, TC2, TC3, and TC4 located at y = 10 mm, aligned at radial distance r = 0, 15, 35 and 55 mm.

determined by cutting the thermocouple after the tests and solid–liquid contact (rewetting) was established. Within
the top sheath thickness was measured. The heat diffusion wetted region the regime changes from film boiling to
of the material around the measuring junction is the same nucleate boiling. Outside this small wetted region, the
as that of the steel plate and the distance of the measuring surface remains dry, partially covered and surrounded by
junction to the surface is precisely known. The calculated rela- a vapour blanket. The wetted region grows fast moving
tive uncertainty of the surface heat flux (qs) is ±5%. The per- radially outward (t = 0.88 s). A white narrow band demar-
formance of this model was validated comparing calculated cates the frontier between wetted and dry zone (rewetting
temperatures with temperature history measured by thermo- front). Sketches demonstrating the first stages of the
couples located 4 mm from top surface at r = 0, 15, 35 and cooling process for Ti = 900°C are shown in Figure 8. In
55 mm, as shown in Figure 7(a). The disagreement found Figure 8(a), an insulating vapour film separates the liquid
between these results was about ±0.15% (Figure 7(b)) at r = from the hot surface. In Figure 8(b), the solid–liquid
0. This results show that quadratic elements of 0.5 × contact was established and a small wetted zone is
0.5 mm² chosen are suitable for the present numerical observed beneath of the jet. Figure 8(c) shows the radial
analysis. advancement outward of the rewetting front where liquid
water droplets are ejected upward at this front. These obser-
vations are in line with Lee et al. [13].
Results and discussion Within this wetted region, the top surface of liquid water
flow appears smooth and shiny (Ti = 900°C; t = 0.88–9.84 s)
Visual analysis
without visible evidence of bubble boiling activity. Analysing
A sequence of snapshots of the jet cooling experiment for external images of the impingement zone, absence of vapour
initial test temperature (Ti) of 600°C and 900°C, illustrates bubbles also was reported by Karwa et al. [14] and Lee et al.
the rewetting process and growth of the wetted zone [13] for water jet cooling of hot steel. They [13,14] argued that
from the centre of the plate, as shown in Figure 8. The Vj and jet subcooling can suppress the bubbly activity.
plate surface initially at 900°C becomes covered by an According to Timm et al. [31] the vapour bubbles formed
insulating vapour blanket after the jet has impinged on its on the hot surface condenses into subcooled liquid. Such
surface (t = 0.08 s). A small dark (wetted) zone is formed bubbles have an average radius of 0.1 mm and a short life-
at t = 0.24 s under the jet, indicating that a stable time of 10−4 s, between growth and collapse.

Figure 7. (a) Comparison between measured and calculated temperature at r = 0, 15, 35 and 55 and 4 mm under surface, for Ti = 900°C; (b) error between
measured and calculated temperature at r = 0 and 4 mm under surface, for Ti = 900°C.
8 H. LEOCADIO AND J. C. PASSOS

Figure 8. Sequence of snapshots from the water jet impingement cooling for Ti = 600°C and 900°C show the rewetting process and growth of the wetted zone
from the centre of the plate. Sketches for Ti = 900°C explain the first stages of the cooling process: (a) a film boiling layer (t = 0.08 s) separates liquid water from hot
surface, (b) solid-liquid contact already has been established (t = 0.24 s) and (c) wetted and dry regions are demarcated by rewetting front (t = 0.88 s) moving
radially outwards.

Even though, the absence of boiling activity in wetted [14]. The cooling tests for Ti = 600°C (see Figure 8) and 750°C
region observed by external images in Figure 8, high-speed also had a similar behaviour observed for Ti = 900°, i.e. with
images from within wetted zone, shown in Figure 2, revealed vapour film formation followed by appearance of dark zone
the formation of a stable vapour film and the growth process and Vrw decreasing as moves radially outwards. However,
of the rewetting area under the jet, followed by vigorous rewetting front moved faster for Ti = 600°C than 900°C. At
boiling activity. Bubbles with average radius of 0.5 mm and t = 9.84 s radial advance was 44 and 31 mm for Ti = 600°C
short lifetime of 10−4 s were visualized growing on surface and 900°C, respectively. This behaviour occurs because a
and condensing into the liquid. Therefore, in Figure 8, even higher Ti increases the amount of energy stored in the plate
though the appearance of absence of bubbles, vigorous and increase the time needed to extract energy, reducing
boiling regime occurs in wetted impingement zone. Vrw. Therefore, higher Ti reduces Vrw. Wang et al. [44] investi-
For Ti = 900°C, the rewetting front velocity (Vrw) decreases gated the local heat transfer characteristics of multi jet
as moves radially outwards. The wetted zone quickly enlarges impinging on hot plate surfaces similar to in the industrial
from t = 0.08 to 0.88 s (0.80 s) with a radial advance of 18 mm scale, with jet velocity ranged between 2.0 and 6.6 m s−1.
and, slowly, from t = 0.88 to 9.84 s (8.96 s) with a radial The results demonstrated that both the hydrodynamic struc-
advance only of 12.5 mm. Agrawal et al. [20] also reported ture and the heat transfer region distribution of multi jet
reduction on Vrw for impinging jet on 800°C-steel surface. impingement cooling were distinct from the single jet case.
This reduction in rewetting front velocity is explained: (i) as Therefore, the present results are valid for single impinging
the water travels radially outwards, it receives heat from the jet.
hot surface reducing its subcooling, i.e. increasing its temp- Figure 9 shows the growth of the wetted diameter for Ti =
erature. With a lower subcooling, its ability to condense 600°C, 750°C and 900°C as function of time. At t < 0.8 s the
vapour bubbles is reduced in the rewetting front edge, and growth rate was unaffected by Ti. However, for t > 0.8 s,
(ii) as the water travels radially outwards its liquid jet layer the growth rate was strongly affected by Ti where a higher
thickness is reduced, becoming much more easier to increase growth rate occurs for a lower Ti. Hatta et al. [17] and
the water temperature. This analysis is in line with Karwa et al. Karwa et al. [14] suggested power law correlation to predict
IRONMAKING & STEELMAKING 9

the rewetting front radius (Rwet) as a function only of time, t:


Rwet = Ctn . Hatta et al. [17] considered 10 mm-circular water
jet at 20°C cooling of a 304 SS plate at 900°C. Karwa et al.
[14] carried out a 3 mm-circular jet impingement cooling of
a 304 SS specimen at 900°C with water jet (2.5 ≤ Vj ≤
10 m s−1; 13 ≤ Tj ≤ 40°C). Karwa et al. [14] found exponent
‘n’ equal to 0.3 ± 0.05 and constant ‘C’ increases with water
jet velocity and subcooling, in the range of 6–14. Hatta
et al. [17] found exponent n ≈ 0.5 and C = 16.7. In both
studies [14,17], the influence of Ti was not taken into
account. Based in the experimental results of wetted diam-
eter shown in Figure 9, an empirical correlation to predict
the Rwet as a function of Ti and time (t) is proposed:
Rwet = (37.9 − 0.03465Ti ) ln (t) − 0.033Ti + 47 (1)
Equation (1) is valid for 600°C ≤ Ti ≤ 900°C, 0.4 s ≤ t ≤ 10 s,
and water jet at 22°C. A comparison between experimental
and calculated values of Rwet is shown in Figure 10. The cor-
relation proposed for Rwet predicts 97% of experimental data
within an error band of 12%. The predicted values have a Figure 10. Comparison of experimental and calculated Rwet with proposed cor-
standard deviation, σ, of ±1.6 mm (95% confidence level). relation Equation (1).
The parameter R-square is 98%.

outwards due to the reduction in water layer thickness, as


explained aforementioned (section 3.1). Figure 12 shows
Cooling curves
the comparison of calculated surface temperatures, Ts, with
Figure 11 shows the measured temperature history by ther- are measured temperatures, Tm, for Ti = 900°C and r = 0.
mocouples 1–5, located 4 mm under top plate surface at When the water jet touches the place (t = 3.2 s), the surface
radial position r = 0, 15, 35, 55 and 75 mm, for test Ti = 900° temperature (Ts = 893°C) is a little lower than the internal
C. The jet touched the hot surface at t = 0 and internal temp- temperature (Tm = 900°C) due its exposition to air. The temp-
erature was Tm = 900°C. At this time, all the cooling curves erature starts to drop at measured position 0.1 s after the jet
coincide and decrease slowly at same cooling rate. Only impingement. After 4.1 s (t = 7.3 s), the surface is Ts = 388°C
about 1.5 s a sharp drop in temperature is observed at r = 0 and internally is Tm = 728°C. This shows the highly nonlinear
and 15 mm, due to the measuring position is 4 mm below temperature profile in the plate just 4 mm from the quench-
top surface. After rewetting has taken place, the heat flux is ing surface. This trend was observed all tests.
enlarged and the temperature drop more evident at position Figure 13 shows isotherms at t = 16.8 s after onset of jet
r = 0 than other positions where hot surface remains dry. cooling process for Ti = 900°C, where a strong thermal nonli-
About 2 s a remarkable increase in the cooling rate occurs nearity inside the plate is observed. At the impingement
at r = 35 mm with the arrival of the rewetting front. The point on surface, at r = 0 and y = 14 mm, the temperature is
same behaviour is observed at position r = 55 and 75 mm. 156°C while at r = 0 and y = 0, the temperature is 624°C.
The intensity of cooling rate reduces as moves radially Figure 14 shows the surface cooling curve and visual
analysis of the first stages of the jet cooling process at

Figure 9. Growth curves of the wetted diameter as a function of time for Ti = Figure 11. Measured internal cooling temperatures at various radial positions
600°C, 750°C and 900°C. for Ti = 900°C.
10 H. LEOCADIO AND J. C. PASSOS

its high temperature. Agrawal et al. [20] reported wetted sur-


faces above 700°C, during the impinging jet at 22°C on 800°C-
steel surface. Others experiments [4,11,13,19] with impinging
jet on hot steel surface also reported rewetting on surface
above critical point.
The rewetting mechanism on such high-temperature sur-
faces was elucidated in authors’ previous work [11,26] exam-
ining the surface conditions by high-speed imagens that
allow direct observation within impingement zone, as
shown in Figures 2 and 3. Since the rewetting temperature
at interface solid–liquid cannot exceed Ttls, the explanation
for such high rewetting surface temperatures is in the
surface condition. Combined with high water subcooling of
80K, which reduces the thickness of the vapour film, rough-
ness can significantly increase both the rewetting tempera-
ture and the heat transfer rate. The asperities protrude
through the thin vapour film, disturbing its continuity and
Figure 12. Comparison between surface and measured-internal temperatures inducing a change from film to transition boiling regime.
at position r = 0 for Ti = 900°C.
Asperities extending into the vapour film interact with
liquid, thereby experiencing a reduction in temperature and
propagate rewetting process on vicinity asperities and
other portions of the surface. Wang et al. [15] and Filipovic
et al. [4] also argued that surface roughness promotes
liquid–solid contact, increasing Trw and reducing trw. Paul
et al. [27], Bui and Dhir [28] and Sinha et al. [24] reported
that roughness, oxidation, and deposition on the plate
surface increase wettability and number of nucleation sites,
causing an earlier collapse of vapour film, enhanced rate of
heat transfer and enlargement of rewetting temperature.
Figure 13. Thermal profile inside the test plate 16.8 s after onset of jet cooling
for Ti = 900°C, from 2D inverse heat conduction model.

Boiling curves
position r = 0 for Ti = 900°C. The cooling starts at Ts = 893°C, t
Figure 15 shows the boiling and HTC curves for Ti = 900°C at
= 0. After 0.11 s, an insulating vapour blanket covers the
position r = 0. Surface heat flux (qs) curve shows that the tran-
surface, where the heat flux is low and the surface tempera-
sition regime (partial areas of the surface are contacted with
ture drops to 890°C. The liquid moves smoothly over the
the liquid water) occurs between 506°C and 893°C. The
vapour blanket. At t = 0.23 s, Ts = 885°C and a very small
maximum heat flux (qmax) of 3 MW m−2 occurs at Ts = 506°C
dark zone under the jet, surrounded by vapour blanket, indi-
and delimits the boundary between the transition and nucle-
cates the surface is wetted and the rewetting already has
ate boiling regimes (bubbly activity). The single phase regime
occurred. At t = 0.38 s, the wetted zone expands radially
begins at Ts ≈ 145°C and qs ≈ 1.2 MW m−², with a clear inflec-
and the surface remains very hot (Ts = 873°C), even after 2 s,
tion and change in the slope of the boiling curve. High quality
when Ts = 574°C beyond by far of critical point of 374°C.
images of 20°C-impinging jet on a 900°C-steel plate, showed
The visible absence of bubbles in wetted region, in spite of
bubbly activity ceasing around Ts = 240°C, in recent work [11].
the high surface temperature, was explained in section 3.1.
According to Ochi et al. [22] and Ishigai et al. [21], the rewet-
ting always occurs for a water jet at below 60°C, regardless of

Figure 14. Surface cooling curve at position r = 0 and visual analysis of the first Figure 15. Heat flux and heat transfer coefficient (HTC) curves as a function of
stages of the jet cooling process for Ti = 900°C. Ts at r = 0 for Ti = 900°C.
IRONMAKING & STEELMAKING 11

In the early stages, both qs and HTC are small due to the
insulation vapour film layer. During the rewetting (transition
regime) occurs the enlargement of surface fractions in
direct contact with the liquid water, increasing the qs and
HTC until qmax is reached and marks the beginning of the
nucleate boiling regime, where the heat flux curve slope
becomes negative. The HTC continues increasing due to the
vapour bubbles number reduction on the surface, which
facilitates the liquid contact with surface. This gradual
reduction bubble population is owing to the surface temp-
erature reduction, which causes a reduction of the number
of active nucleation sites. At the beginning of the single-
phase regime, the slope of HTC keeps increasing as Ts
decreases. This behaviour can be explained by examining
the thermal boundary layer thickness. As Ts decreases, the
thermal boundary layer thickness diminishes allowing larger
HTC [45]. Sikdar et al. [9] reported HTC increases with a
decrease in temperature from strip surface experiences on
the run out table.
Figure 16 shows the effect of Ti on the boiling curves. For Figure 17. Effect of Ti on the HTC, for Ti = 600, 750 and 900°C, at position r = 0.
Ti = 600°C, 750°C and 900°C, qmax = 2.1, 2.5 and 3.0 MW m−² at
Ts = 340°C, 365°C and 506°C, respectively. Thus, increasing Ti All curves shows an increase, with different slopes, as Ts is
will occur a rise in qmax. After qmax, all boiling curves decay reduced until Ts ≈ 300°C where the curves meet around
overlapped. This behaviour shows that the single-phase and 8 kW m−² K−1 and remain together through single-phase
nucleate boiling regimes are unaffected by Ti and Ts. After forced convection. From Ts ≈ 160°C all curves experience a
the maximum heat flux has been reached, the heat transfer sudden enlargement in their slope, reaching values close to
occurs through the fully developed nucleate boiling regime. 18 kW m−²°C. Therefore, HTC is affected by Ts and Ti constant
A review of available fully developed nucleate boiling data, values of HTC must be avoided in cooling control models in
in jet impingement boiling, by Wolf et al. [46] revealed that the range of 200°C < Ts < 900°C.
conditions depend strongly on fluid used. For fully developed The effect of Ti on the boiling curves for Ti = 600°C and
nucleate boiling (FNB) of water jet, surface heat flux can be 900°C, at positions r = 0, 15, 35, 55 mm, is shown in Figures
correlated by empirical equation: qFNB = C.DTsat n
, where qFNB 18 and 19, respectively. After rewetting has taken place, the
and ΔTsat (surface superheat) have units of W m−² and °C, heat flux is enlarged more evident at position r = 0 than
respectively. ‘C’ and ‘n’ depends on the fluid and type of other positions where hot surface remains dry.
jet. Since the type of jet and fluid used are fixed in present As rewetting front arrives at each radial position, the peaks
study, the surface heat flux depends only on the surface of qmax occur in the curves. The peaks decrease as they move
superheat, so the slope of the boiling curve will have the away from r = 0. The reduction in the peaks is caused by the
same slope. For all tests, the boiling curves exhibit the heating of water as it moves radially outwards, reducing its
same inflection point around Ts ≈ 140°C demarking heat removal capacity, as explained in section 3.1. This
the onset of the single-phase regime. Figure 17 shows the trend was observed in all tests. The time spent to reach
effect of Ti on the HTC at r = 0 for Ti = 600, 750 and 900°C. qmax was 2.8 and 3.6 s for Ti = 600°C and 900°C, respectively.

Figure 16. Effect of Ti on the heat flux for Ti = 600, 750 and 900°C, at position Figure 18. Surface heat flux curves as a function of time for Ti = 600°C at pos-
r = 0. itions r = 0, 15, 35, 55 mm.
12 H. LEOCADIO AND J. C. PASSOS

Figure 19. Surface heat flux curves as a function of time for Ti = 900°C at pos-
itions r = 0, 15, 35, 55 mm.

Figure 20. Comparison between experimental and calculated values of HTC


As aforementioned, higher Ti means higher qmax and, conse- with equation (2).
quently, higher rewetting delay time.
For Ti = 600°C (Figure 18), the heat flux peaks at position
r = 0, 15, 35 and 55 mm reach a value of 2.15, 1.44, 1.16 and properties for a hot steel strip mill, where surface temperature
0.80 MW m−² at t = 9.09, 9.89, 14.1 and 28.9 s, respectively. and heat flux are typically very large and acceptable cooling
The time interval between the peaks increases as the rewet- times are relatively short. In order to illustrate the behaviour
ting front moves radially outward: 0.8, 4.2 and 14.8 s due to of the impinging jet employed during accelerated cooling in
reduction in the rewetting front velocity. The qmax at r = 0 the hot rolling process, Figure 21 shows a hot strip mill at 900°
was 49% higher than at r = 15 mm and 168% higher than at C moving at 6 m s−1 (a) and 10 m s−1 (b) through the first of
r = 55 mm, demonstrating the higher cooling efficiency in several planar jets (water curtain) on the runout table at hot
the impingement zone. For Ti = 900°C (Figure 19), the heat strip mill line of Usiminas Steel.
flux peaks at position r = 0, 15, 35 and 55 mm were 3.0, 2.2, Figure 22 shows a 20-mm-thick 850°C-steel plate moving
1.7 and 1.0 MW m−², with time intervals between peaks of at 18 m min−1 throughout arrays of impinging circular
11.0, 15.3, 29.6 s, respectively. The intervals between the water jets in the hot rolling mill pilot plant at R & D Center
peaks were larger than those for Ti = 600°C. Again, these of Usiminas. In both cases (Figures 21 and 22) the dark
results show that high Ti slows down the rewetting front vel- (wetted) zone is clearly visible beneath the impinging jet.
ocity. Agrawal et al. [20] reported similar trend for surface An outer white zone (vapour blanket) surrounds the wetted
heat flux, during the impinging jet at 22°C on 800°C-steel zone, similar to cooling test shown in Figure 8. Owing to
surface, and that maximum heat flux decreases as the pos- the surface movement, the liquid water layer stretches in
ition moves away from the stagnation point (r = 0). Wang the direction of motion, making the length of wetted zone
et al. [10] reported heat flux peaks decrease as they move stretches more downstream than upstream, acquiring a
away from r = 0, for a 4 mm-circular jet impinging (10°C) at non-symmetric shape. For round impinging jet on plate
4.46 m s−1 on top of the hot steel plate (670°C), with similar moving at 18 m min−1 (see Figure 22) the wetted zone size
heat fluxes values. Based in the experimental results of HTC is around 8 Dj long. For planar jet on strip at 6 m s−1 (see
values (Figure 17), the following empirical correlation to Figure 21(b)) the wetted zone size is around 20 planar jet
predict HTC in a single circular impinging water jet, as a func- thicknesses long. These results are larger than those reported
tion of Ts and Ti, is here proposed: in some studies [5–7]. Wiskel et al. [6] also reported obser-
vations of elliptical shape in hot strip mill line, but, of only
(Ts − 240)1.29 2 Dj long. Other studies [5,7] considered circular shapes of
HTC = 8368 − 13179 (2)
(Ti − 240)1.38 only 2–3 Dj. Thus, static circular wetted zone must be
avoided in cooling models. For planar jet on strip at
Equation (2) is valid for 600°C ≤ Ti ≤ 900°C, and water jet at
10 m s−1 (see Figure 21(a)) the wetted zone is almost invisible.
22°C. A comparison between experimental and calculated
This phenomenon is explained since higher strip speed
values of HTC is shown in Figure 20. The correlation proposed
means less resident time of a specific point on the strip
for HTC predicts 92% of experimental data within an error
surface below the impingement zone, and consequently
band of 5%. The predicted values have a standard deviation,
less time for rewetting to start. Therefore, the effect of
σ, of ±107 W m−².°C (95% confidence level). The parameter R-
surface motion has to be considered in the cooling control,
square is 98%.
since effective cooling occurs within wetted region which
enlarges as strip speed reduces. Higher velocities can result
in a low cooling efficient due to small wetted zone. The
Heat transfer during hot strip mill cooling
strip speed limit can be estimated by rewetting delay corre-
Temperature cooling control at the runout table is one of the lation [11] as function of Vj, Ti and Tj. Furthermore, rewetting
most critical processes for ensuring the required mechanical delay time (trw) reduces with the decreasing of Ti, as
IRONMAKING & STEELMAKING 13

Figure 21. Photographs of the dark zone formed on 4 mm-thick hot steel strip mill surface at 900°C moving at (a) 10 m s−1 and (b) 6 m s−1 crossing the first of
various impinging water curtain jets (planar jet) on the runout table during its cooling process.

occurred. Rewetting (liquid–solid contact) takes place on


surface above 880°C, beyond of critical point of water.
The rewetting delay time (trw) enlarges as initial tempera-
ture (Ti) rises. Higher Ti delays the onset of the rewetting
and reduces the rewetting front velocity.
. An increase in Ti causes a rise in the maximum heat flux
(qmax). High heat fluxes occur within wetted impingement
zone at r = 0 and decrease as move radially outwards.
. The heat transfer coefficient (HTC) is affected by Ti and
surface temperature (Ts). HTC enlarges as Ts decreases
during the jet cooling. Correlations for HTC and rewetting
Figure 22. Photograph of the wetted (dark) zone formed during accelerated
cooling of a 20 mm-thick steel plate at 850°C moving at 18 m min−1 going front radius (Rwet) have been proposed.
through arrays of impinging round water jets in a hot rolling mill pilot plant. . Visible proofs showed a non-symmetric effective cooling
zone (wetted zone) stretched more downstream due to
the surface motion. For a strip at 6 m s−1, wetted zone
demonstrated in section 3.2, and then larger wetted zone will
size ≈ 20 planar jet thicknesses, while at 10 m s−1 is
appear as strip surface temperature is reduced on run-out
almost nonexistent. Low strip velocities produce larger
table. Therefore, knowledge of the trw, Trw and wetted zone
wetted zone increasing the heat removal capacity.
growth are crucial in hot steel rolling cooling control. . Results will contribute for enhancement and designing of
effective industrial ultra-fast cooling systems.
. Heat transfer behaviour in jet impingement boiling on
Conclusions
high-temperature steel surface has revealed to be highly
The heat transfer during the cooling of 600–900°C- complex, requiring further research, with multi jets on
steel plate by an impinging water jet was characterized hot moving surfaces, employing high-speed cameras to
using an experimental apparatus and an inverse heat provide a direct observation of boiling phenomena
conduction model: within impingement zones, interacting with adjacent
jets. This experiment will provide knowledge closer to
. An effective cooling zone (wetted zone), with high heat the reality of the industrial line being useful in cooling
flux, appears and grows only after the rewetting has system design.
14 H. LEOCADIO AND J. C. PASSOS

Acknowledgements [22] Ochi T, NakanishI S, Kaji M, et al. Cooling of a hot plate with an
impinging circular water jet. In: T Veziroglu, A Bergles,
Authors acknowledge the support from Usiminas Steel. Amsterdam: Elsevier Science Publishers B.V.; 1984. p. 671–681.
[23] Liu Z, Wang J. Study on film boiling heat transfer for water jet
impinging on high temperature flat plate. Int J Heat Mass Transf.
Disclosure statement 2001;44:2475–2481.
No potential conflict of interest was reported by the author(s). [24] Sinha J, Hochreiter L, Cheung F. Effects of surface roughness, oxi-
dation level, and liquid subcooling on the minimum film boiling
temperature. Exp Heat Transfer: J Thermal Energy Generation,
ORCID Transport, Storage, Conversion. 2003;23(1):45–60.
[25] Lienhard JH, Shamsundar N, Biney PO. Spinodal lines and equations
H. Leocadio http://orcid.org/0000-0002-7851-6950 of state: a review. Nucl Eng Des. 1986;95:297–314. https://doi.org/
J. C. Passos http://orcid.org/0000-0001-7922-4039 10.1016/0029-5493(86)90056-7.
[26] Leocadio H, Geld Cvd, Passos JC. Degassing, boiling and rewetting
in free surface jet quenching. 9th World Conference on experimen-
References
tal heat transfer, fluid Mechanics and Thermodynamics, Iguazu
[1] Hasan MN, Monde M, Mitsutake Y. Homogeneous nucleation boiling Falls, Brazil. 2017.
during jet impingement quench of hot surfaces above thermodyn- [27] Paul G, Das P, Manna I. Assessment of the process of boiling heat
amic limiting temperature. Int J Heat Mass Transf. 2011;54:2837–2843. transfer during rewetting of a vertical tube bottom flooded by
[2] Lee J, Samanta S, Steeper M. Review of accelerated cooling of steel alumina nanofluid. Int J Heat Mass Transf. 2016;94:390–402.
plate. Ironmaking Steelmaking. 2015;42(4): 268–273. [28] Bui TD, Dhir VK. Transition boiling heat transfer on a vertical surface.
[3] Agrawal C. Surface quenching by jet impingement – a review. Steel J Heat Transfer. 1985;107:756–763.
Res Int. 2018;1800285: 1–22. https://doi.org/10.1002/srin.201800285. [29] Prieto M, Ruiz L, Menendez J. Thermal performance of numerical
[4] Filipovic J, Incropera FP, Viskanta R. Rewetting temperatures and vel- model of hot strip mill runout table. Ironmaking Steelmaking.
ocity in a quenching experiment. Exp Heat Transf. 1995;8(4):257–270. 2001;28:474–480.
[5] Cox SD, Hardy SJ, Parker DJ. Influence of runout table operation [30] Zumbrunnen DA, Incropera F, Viskanta R. Method and apparatus for
setup on hot strip quality, subject to initial strip condition: heat measuring heat transfer distributions on moving and stationary
transfer issues. Ironmaking Steelmaking. 2001;28(5):363–372. plates cooled by a planar liquid jet. Exp Therm Fluid Sci. 1990;3
https://doi.org/10.1179/irs.2001.28.5.363. (2):202–213.
[6] Wiskel JB, Deng H, Jefferies C, et al. Infrared thermography of TMCP [31] Timm W, Weinzierl K, Leipertz A. Heat transfer in subcooled jet
microalloyed steel skelp at upcoiler and its application in quantifying impingement boiling at high wall temperatures. Int J Heat Mass
laminar jet/skelp interaction. Ironmaking Steelmaking. 2011;38(1):35–44. Transf. 2003;46:1385–1393.
[7] Filipovic J, Viskanta R, lncropera F. Cooling of a moving steel strip by [32] Sabioni ACS, Ramos RPB, Ji V, et al. About the role of chromium and
an array of round jets. Steel Res. 1994;65(12):541–547. https://doi. oxygen ion diffusion on the growth mechanism of oxidation films
org/10.1002/srin.199401210. of the AISI 304 austenitic stainless steel. Oxid Met. 2012;78:211–220.
[8] Lee PJ, Raudensky M, Horsky J. Development of accelerated cooling [33] Incropera FP, DeWitt D, Bergman TL, et al. Fundamentals of heat
for new plate mill. Ironmak Steelmak. 2013;40:598–604. and mass transfer. 7th ed. Wiley; 2011.
[9] Sikdar S, Mukhopadhyay A. Numerical determination of heat trans- [34] Horsky J, Hrabovsky J, Raudensky M. Impact of the oxide scale on
fer coefficient for boiling phenomenon at runout table of hot strip spray cooling intensity. 10th International Conference on Heat
mill. Ironmaking Steelmaking. 2004;31(6):495–502. Transfer; Fluid Mechanics and Thermodynamics, Orlando. 2014.
[10] Wang B, Lin D, Xie Q, et al. Heat transfer characteristics during jet [35] J. Taylor, An Introduction to error analysis – the study of uncertain-
impingement on a high-temperature plate surface. Int J Heat ties in physical measurements, 2nd ed. University of Colorado, 1996.
Mass Transf. 2016;101:902–910. [36] Stevens J, Webb BW. Measurements of flow structure in the radial
[11] Leocadio H, van der Geld CWM, Passos JC. Rewetting and boiling in layer of impinging free-surface liquid jets. Int J Heat Mass Transf.
jet impingement on high temperature steel surface. Phys Fluids. 1993;36(15):3751–3758.
2018;30:122102. https://doi.org/10.1063/1.5054870. [37] Stevens J, Webb BW. Measurements of flow structure in the stagna-
[12] Carbajo JJ. A study on the rewetting temperature. Nucl Eng Des. tion zone of impinging free-surface liquid jets. Int J Heat Mass
1985;84:21–52. Transf. 1993;36(17):4283–4286.
[13] Lee SG, Kaviany M, Kim C, et al. Quasi-steady front in quench sub- [38] Stevens J, Webb BW. Measurements of the free surface flow struc-
cooled-jet impingement boiling: experiment and analysis. Int J ture under an impinging free liquid jet. J Heat Transfer. 1992;114
Heat Mass Transf. 2017;113:622–634. (1):79–84.
[14] Karwa N, Stephan P. Experimental investigation of free-surface jet [39] Bhatt N, Raj R, Varshney P, et al. Enhancement of heat transfer rate
impingement quenching process. Int J Heat Mass Transf. of high mass flux spray cooling by ethanol-water and ethanol-
2013;64:1118–1126. tween20-water solution at very high initial surface temperature.
[15] Wang Z, Zhong M, Deng J. Experimental investigation on the tran- Int J Heat Mass Transf. 2017;110:330–347.
sient film boiling heat transfer during quenching of FeCrAl. Ann [40] Ravikumar SV, Jha JM, Sarkar I, et al. Ultrafast cooling of medium
Nucl Energy. 2021;150:107842. carbon steel strip by air atomised water sprays with dissolved addi-
[16] Takrouri K, Luxat J, Hamed M. Experimental investigation of quench tives. Ironmaking Steelmaking. 2014;41(7):529–538.
and re-wetting temperatures of hot horizontal tubes well above the [41] Ravikumar SV, Jha JM, Tiara AM. Experimental investigation of air-
limiting temperature for solid–liquid contact. Nucl Eng Des. atomized spray with aqueous polymer additive for high heat flux
2017;311:167–183. applications. Int J Heat Mass Transf. 2014;72:362–377.
[17] Hatta N, Kokado J, Hanasaki K. Numerical analysis of cooling charac- [42] Karwa N, Schmidt L, Stephan P. Hydrodynamics of quenching with
teristics for water bar. Trans Iron Steel Inst Jpn. 1983;23:555–564. impinging free-surface jet. Int J Heat Mass Transf. 2012;55:3677–
https://doi.org/10.2355/isijinternational1966.23.555. 3685.
[18] Woodfield PL, Monde M, Mozumder AK. Observations of high temp- [43] Trujillo DM, Busby HR. Practical inverse analysis in engineering. Boca
erature impinging-jet boiling phenomena. Int J Heat Mass Transf. Raton: CRC Press; 1997; https://doi.org/10.1201/9780203710951.
2005;48(10):2032–2041. [44] Wang B, Lin B, Zhang B, et al. Local heat transfer characteristics of
[19] Xu F, Gadala M. Heat transfer behavior in the impingement zone multi Jet impingement on high temperature plate surfaces. ISIJ
under circular water jet. Int J Heat Mass Transf. 2006;49:3785–3799. Int. 2018;58(1):132–139.
[20] Agrawal C, Kumar R, Gupta A. Rewetting and maximum surface heat [45] Webb BW, Ma CF. Single-phase jet impingement heat transfer. In:
flux during quenching of hot surface by round water jet impinge- Advances in heat transfer, vol. 26. San Diego: Academic Press;
ment. Int J Heat Mass Transf. 2012;55:4772–4782. 1995. p. 105–217.
[21] Ishigai S, Nakanishi S, Ochi T. Boiling heat transfer for a plane water [46] Wolf DH, Incropera FP, Viskanta R. Jet impingement boiling. In:
jet impinging on a hot surface). Proceedings 6th Int. heat transfer Advances in heat transfer, vol. 23. San Diego: Academic Press;
Conference; Toronto, Canada. 1978. DOI: 10.1615/IHTC6.860. 1993. p. 1–132.

You might also like