Sie sind auf Seite 1von 12

Experimental Thermal and Fluid Science 101 (2019) 251–262

Contents lists available at ScienceDirect

Experimental Thermal and Fluid Science


journal homepage: www.elsevier.com/locate/etfs

Dynamics of single droplet impact on cylindrically-curved superheated T


surfaces

R. Simhadri Rajesh, P.T. Naveen, K. Krishnakumar, S. Kumar Ranjith
Micro/nanofluidics Research Laboratory, Department of Mechanical Engineering, College of Engineering Trivandrum (Government of Kerala), Thiruvananthapuram,
Kerala 695016, India

A R T I C LE I N FO A B S T R A C T

Keywords: Liquid droplet impact on superheated symmetric surfaces such as flat and spherical geometries has been ex-
Drop impact tensively investigated in the past. However, the effect of a superheated asymmetric curvature on the collision
Symmetry-breaking dynamics remains unknown. In this study, experiments are performed to elucidate the thermo-hydrodynamic
Residence time behavior of milli-metric water droplets impacting on superheated cylindrical surfaces having convex and con-
Spread factor
cave profiles. These geometries offer an asymmetric zone for droplets to expand and retract. Mono-dispersed
Cylindrically curved surface
water droplets are impinged on convex and concave surfaces maintained in the temperature range of
125–290 °C. The impact Weber number is varied between 5 and 65 and droplet evolution is visualized using the
high-speed imaging technique. It is observed that, droplet morphology and spreading characteristics are entirely
different for convex and concave surfaces compared to a flat substrate due to momentum anisotropy. The
maximum spread factor and residence time of the water droplet are significantly influenced by the impact Weber
number while the surface superheat has negligible effect. Importantly, the asymmetric distribution of mo-
mentum facilitated 16% and 24% reduction in contact time for convex and concave surfaces respectively. The
preferential extension of fluid in the azimuthal direction assisted by the gravity causes greater spreading of
droplets on a convex surface. In contrary, these conditions enforced a reduction in contact time for droplets on a
concave topography.

1. Introduction heat transfer reaches its minimum value and impact dynamics of drops
are drastically altered [11,12].
Droplet impact on solid surfaces is an important phenomenon with The single droplet behavior on hot surfaces is a very important
regard to numerous industrial applications. For instance, during spray- phenomenon, since the available heat transfer area is dependent on the
coating, crop spraying, painting and printing processes the droplets droplet spreading characteristics and duration of contact with the sur-
collide on surfaces at room temperature [1–5]. Wherein, spray-cooling faces [2]. In general, a droplet with initial diameter d 0 hits a super-
of metals and electronics devices, re-flooding of nuclear fuel rods, fuel heated surface and spreads to a maximum diameter dmax . Thereafter, a
droplet impact on combustion chamber walls and fire suppression etc. thin vapor layer forms between the drop and the surface and surface
are examples of drops impacting on hot surfaces [6–10]. Indeed, proper tension forces retract the elongated drop back into spherical shape.
contact between the drop and hot surface and instantaneous formation Subsequently, the radial momentum of the fluid is converted to vertical
of vapor layer have significant influence on the outcome of these pro- momentum and the droplet jump upwards after a residence time, tr .
cesses. Generally, the impact conditions and difference in saturation Two major non-dimensional parameters which dictates the physics of
temperature of liquid and temperature of the superheated surface (TS ) the problem is impact Weber number (We) and maximum spread factor
decide the type of boiling, water vapor creation and concurrent heat (β ). The ratio of inertia forces to surface tension forces is expressed as
transfer effects. At a certain elevated temperature known as Leidenfrost Weber number We = ρvd2 d 0/ σ , where ρ is the density, σ is the surface
temperature (TL ), a thin vapor layer is formed between the sessile drop tension and vd is the droplet velocity. Droplets after impact on the hot
and the solid surface and droplet levitates subsequently on the surface surfaces may spread, splash or rebound depending on the impact ve-
[11]. The formation of this insulating vapor cushion suppress the heat locity, surface properties, liquid properties and surface superheat. An-
flow from the surface to the drop. When surface temperature TS ≈ TL the other important control parameter is residence time (tr ∼ ρd 03/ σ )


Corresponding author.
E-mail address: ranjith@cet.ac.in (S. Kumar Ranjith).

https://doi.org/10.1016/j.expthermflusci.2018.10.011
Received 25 March 2017; Received in revised form 25 May 2018; Accepted 8 October 2018
Available online 15 October 2018
0894-1777/ © 2018 Elsevier Inc. All rights reserved.
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

[13,14], the duration which a droplet remains in contact with the solid experimentally and numerically the impact behavior of water, iso-
surface. Most of the droplet impact studies on hot surfaces were focused propyl alcohol and acetone micro-liter droplets on a heated brass
on the maximum spread factor β = dmax / d 0 [1,4,15] and non-dimen- spherical particle. The effect of Weber number and surface temperature
sional residence time, τr∗ = tr / ρd 03/ σ [15,16] of the droplet in the film on the droplet maximum spread factor and contact time of the droplet
boiling regime. were examined. In the film boiling regime, surface wetting was pre-
Followed by Worthington’s landmark experiment [17], numerable vented by a thin vapor film at the contact area which led to a smaller
investigations were performed to understand the complex dynamics of temperature drop at the surface.
single droplet impact on solid surfaces [1,3,4,18–23]. In particular, Most importantly, Liu et al. [22] conducted experiments to study
studies involving droplets collision on superheated surfaces assumed bouncing of liquid droplets on superhydrophobic cylindrical surfaces
importance due to its engineering implications and were reported in kept at normal temperature. The droplet initially extends symme-
detail using high-speed imaging techniques [5,12,14,24–26]. Note that, trically, but the symmetry breaks with retraction in axial direction
most of these studies were focused on droplet impingement on flat solid while the droplet is in spreading phase in the azimuthal direction on
surfaces. Major contributing parameters on heat transfer enhancement convex surface and the drop rebounds from the surface. This phenom-
are surface temperature, surface morphology, surface wettability, drop enon is also observed in the concave surface, but the retraction is along
size, liquid boiling point, impact velocity, contact area and residence the azimuthal direction when the liquid is in spreading phase in the
time. Wachters et al. conducted water droplet impingement experi- axial direction. Further, it was reported that, the asymmetry in boun-
ments on a heated gold surface at 400 °C and found that the droplets cing leads to approximately 40% reduction in contact time.
disintegrate into smaller droplets after rebound from the surface when It is noteworthy that, most of the superheated droplet impact dy-
30<We<80 and beyond which droplet disintegrates into smaller dro- namics were studied on plane flat surfaces. In a practical scenario,
plets [27]. Hatta et al.[28], studied collision dynamics of water droplets importance of liquid droplet impact is not only relevant to flat surfaces,
on a heated surface (TS = 500 °C) above the Leidenfrost temperature and but also to cylindrical surfaces. For instance, spray cooling of metal rods
proposed that the droplet after impact ruptures into secondary droplets in metal processing industries, cooling of nuclear reactor fuel rods and
around a critical We = 50 . It was also reported that, the maximum fire suppression of cylindrical metal columns in buildings etc. involves
spreading factor β [29,30] and angle of incidence [31] of drops are profiles of circular shape. In these situations, the geometry of the sur-
found to be influencing the heat transfer process. Further, the impact faces is asymmetrical, since it is curved in the azimuthal direction and
We and surface superheat are the crucial factors influencing the rate of flat along the axial direction. The effect of surface curvature on the
evaporation and concurrent levitation of drops and subsequent recoil droplet dynamics is not studied in detail, although there are some
cycles [32–34]. Cossali et al.[35,36] experimentally studied secondary studies on spherical surfaces [5,41] and non heated cylindrical surfaces
atomization of a liquid drop impacted on heated walls and reported that [22,38,40]. However, the effect of curvature of the surface in case of a
the secondary atomization occurred due to thermal effects. For low superheated cylindrical surface on the droplet behavior still remains
viscous liquids, a central liquid jet was observed in the film boiling uncovered. The objective of the present work is to study the effect of
regime, but disappeared for highly viscous fluids. It was explained that surface curvature on the dynamics of the impinging water droplets. The
the generation of pressure wave at the impact point due to rapid for- experiments are carried out at different superheat temperatures from
mation of a central vapor bubble is the main reason for the liquid jet 125 °C to 290 °C using high-speed photography. Since, the current study
formation. Further, Liang et al. [16] observed a central liquid jet using focus on the droplet bouncing and a high Weber number facilitates
NaCl solution owing to the bubble entrainment with violent nucleation. secondary droplet generation, the impact Weber number is limited
Tran et al.[14] experimentally studied the effect of micro-structures on between 5 and 65. Quantitative study of the effect of the surface cur-
the τr∗ and β and noticed that it is not influenced by droplet velocity. In vature ratio, surface superheat and droplet impact Weber number on
contrast with a smooth surface a structured surface creates a central the maximum spread factor and residence time of the droplet in the film
liquid jet. In addition, the role of wettability of substrate was examined boiling regime are carried out and compared with a standard flat sur-
to understand its effect on droplet dynamics. For a hot hydrophilic face.
surface the residence time increases drastically [37] while a super-hy- Organization of the paper is the following manner. The experi-
drophobic surface at laboratory condition exhibited a breakdown in the mental procedure of high-speed imaging technique and image analysis
theoretical limit of contact time [19]. Moreover, droplet contact time is is described in Section 2. Detailed results and discussion of droplet
an important parameter influencing heat transfer and it is obeserved collision on convex surfaces is presented in Section 3 while that of
that, in the film boiling regime it is independent of impact velocity concave surfaces is given in Section 4. A qualitative and quantitative
[16,27,28,34]. All these studies mainly focused on the drop interaction comparison of single droplet impact on flat, convex and concave pro-
with flat solid substrate, however in several practical applications the files is discussed in Section 5 before we conclude.
impact dynamics of curved surfaces is important as well [5,22].
The curvature of impinging surface plays vital role in the spreading, 2. Experimental details
disintegration and formation of secondary droplets. Different groups
investigated the drop impact on cylindrical [38–40] as well as on This section details about the experimental setup, test surface pre-
spherical surfaces [5,41–43] under cold and hot conditions. Hung and paration and quantification of observations using image analysis.
Yao [38] performed water droplet impact experiments on cylindrical
wires. The droplet diameter varied between 110 μm and 680 μm and 2.1. System description and procedure
observed that the micro-droplets disintegrated when wire diameter is
small or the impact velocity is high. In contrary, dripping of the dro- Since the droplet impact sustain only for a duration of a few milli-
plets from the bottom side of the wires was noticed when the wire seconds, high-speed imaging technique is employed to capture the se-
diameter is large or the impact velocity is low. Shen et al. [40] con- quence of droplet deformation. The experimental setup mainly consists
ducted high-speed imaging experiments on the liquid droplet behavior of a heater module with a temperature controller, test surfaces, a dro-
on dry and wetted cylindrical surfaces. The impact behavior differed plet dispensing system, a high-speed camera, a back light for illumi-
greatly when the cylinder-drop curvature ratio (relative strength of nation and a workstation for data storage and analysis. All these com-
droplet to curvature diameter) was altered from 0.091 to 2.395. The ponents are firmly attached to an optical breadboard for stability as
effect of surface curvature ratio on maximum spreading factor is neg- shown in Fig. 1. The heater module consists of a custom built heater
ligible when the curvature ratio is less than 0.5. Evans et al. [5] studied block and a temperature control system for achieving isothermal con-
ditions. The outer body of the heater block is made up of mild steel and

252
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

Fig. 1. Schematic diagram of the experimental setup for visualizing droplet impact.

calibrated 1.5 mm diameter K-type thermocouple probe (range


0–1000 °C) is tightly fit into a hole drilled on the test surfaces, 2 mm
below the point of droplet impact for measuring the surface tempera-
ture. It is connected to the temperature controller/display for control-
ling the surface temperature.
In the current study, the droplets are allowed to collide on surfaces
with three different geometries; (i) flat, (ii) convex and (iii) concave. A
convex profile is a semi-cylindrical surface protruding out and a concave
profile is a semi-cylindrical groove carved out from a flat surface as
shown in Fig. 2(a)–(d). Both the convex/concave surfaces are having
two distinct diameters DS ≈ 12 mm and 8 nm , as given in Fig. 2(a) and
(c). These profiles are fabricated from a pure aluminium block em-
ploying wire cut EDM. Further, SEM images of the surface texture of the
exposed surface after burning is shown in Fig. 2(e). The sessile drop
contact angle is measured using a contact angle goniometer using an
optical system to capture the profile of a water on the solid substrate.
The image of sessile micro drop resting on the Aluminium surface at
five different places are taken using silhouette technique. The images
are used for measurement of contact angle using an image analysis
software and is averaged. Typically metal surfaces are hydrophilic in
nature, however the test specimen made from wire-cut EDM exhibited a
high contact angle of θc = 125.06° ± 3 as seen in Fig. 2(f) with an
average contact angle hysteresis of 6°. Here, the surfaces are hydro-
phobic due the surface modification obtained through wire cut EDM
process. It was reported that, due to the ex-foliation characteristics of
aluminium, high contact angles (156 °C) can be achieved through wire
EDM [44]. The presence of microns sized roughness on the surface as
evident in Fig. 2(e) facilitates the entrapment of air and subsequent rise
in hydrophobicity.
The test liquid chosen for realizing drops is distilled water.
Throughout the experiments an initial droplet temperature of
Fig. 2. Geometric details of cylindrically curved aluminium test surface. Front Tf = 28 ± 2 °C is ensured and the radiation effect from the surface is
and top views of the specimen having (a & b) convex and (c & d) concave
neglected for consistency. Properties of the water at Tf = 28 °C is taken
shapes. SEM image (Nova nanoSEM NPE206) of the surface texture with (e)
as, thermal conductivity λ = 0.6098 W/mK , density ρ = 998 kg/m3 ,
800× magnification. (f) Image of a static water drop on top of the flat test
surface tension σ = 71.516 N/m and viscosity μ = 8.6466 × 10 4 N. s/m2 .
section with contact angle 125.06°.
Droplet dispensing system comprises of an xyz-stage arm fitted with a
fluid reservoir having a 26G × 1/2″ hypodermic needle attached (outer
a heater element in enclosed in it. A 150 Watt capacity nichrome-coil diameter: 0.46 ± 0.0064 mm, inner diameter: 0.260 ± 0.019 mm). The
embedded stainless steel plate (100 mm × 100 mm) is used for heating water droplets freely fall under the influence of gravity through the
the surfaces. It is placed centrally in the heater block with the bottom needle tip. Mono-disperse droplets are generated with the help of a flow
side insulated with phenolic tile and glass wool. Leveling screws are control valve such that the droplet size is a sole function of needle
attached to the base of the heater block and the temperature of the test diameter. The images of droplet dynamics are captured using a Hi-
surfaces is controlled using a temperature controller (Selec make). A

253
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

Fig. 3. Schematic of droplet spreading in


azimuthal direction for (a) convex and (b)
concave surfaces with curvature diameter,
Ds .

Fig. 4. Droplet impingement on a convex surface having curvature ratio, δ = 0.29. (a)–(e) Behavior of Liedenfrost drops on solid surface for different temperatures at
We = 16.48 , images are captured after a time duration of (a) 71 ms (b) 21 ms, (c) 4.75 ms, (d) 10.5 ms and (e) 12 ms after the collision.

Fig. 5. Image sequences of droplet time evolution on a convex surface with δ = 0.29, TS = 275 °C and We = 5.41. Shape change of droplet with respect to time viewed
along (a)–(e) azimuthal and (f)–(j) axial directions. Note that, both the views are captured for different drops under identical conditions.

Spec2 high-speed camera (Fastec Imaging, USA) equipped with a temperature reached the desired preset value under steady state con-
Tamron 90 mm f/2.8 macro lens. The high-speed camera continuously ditions. The experiments are conducted for flat, convex and concave
sweeps the images at 4000 fps with a resolution of 362 × 320 pixels for profiles sequentially. The high-speed camera is aligned horizontally and
a shutter speed of 200 μs . An LED light with a diffuser plate is used for the impact dynamics are captured in the azimuthal direction for the
back illumination and shadow images of the droplets are recorded. The convex and concave surface profiles. The impact velocity of the droplet
experiments are conducted at atmospheric pressure and temperature in is varied by changing the droplet release height h, above the test sur-
lab conditions. Test surfaces are placed on the heater and the surface face.
temperature (TS ) is varied between 125 °C and 290 °C using a tem- Quantification of droplet deformation parameters are performed by
perature controller. The average surface temperature at the point of processing images obtained from the high-speed camera which en-
droplet impact is measured alternatively using a calibrated digital compasses both spacial and temporal information. Analysis of the
thermometer (range 0–300 °C) over an arc length of ± d 0 from the point images is carried out using ImageJ (V 1.50b) free software from NIH,
of impact. Due to the high conductivity of aluminium the difference in U.S.A. The calibration of the image analysis is performed by setting the
temperature between the point of droplet impact and thermocouple pixel count of the captured images in the scaling module of the soft-
junction for a concave, flat and convex surfaces is measured less than ware. Thereafter, pixel analysis is carried out to find the droplet dia-
2 °C, 1.75 °C and 1.5 °C respectively. The shadow images of the droplet meter. The error in pixel measurement is ± 1 pixel with a size of
impacts are captured using high-speed camera after the surface ± 0.047 mm . The initial droplet diameter is calculated at the time of

254
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

Fig. 6. Shape evolution of water droplet impinged on convex surfaces kept at TS = 275 °C for different curvature ratios and Weber numbers.

release by averaging 50 trials from different angles to assure the 3. Droplet impact dynamics on superheated convex surfaces
sphericity of drop and is found to be d 0 = 2.33 mm . Further, the impact
velocity vd , and residence time tr , of the impinging water droplets are In this section, visualization of droplet collision on semi-cylindrical
estimated from the simultaneous space and time data. The drop impact convex surfaces is described. The geometry of these surfaces is asym-
velocity is calculated by tracking the drop centroid location between metric, which are curved in the azimuthal direction and flat along the
two images of 0.25 ms time spacing just before impact. Similarly, axial direction. Since, the curvature of surface is in the azimuthal di-
average impact velocities are estimated for different droplet release rection, the dynamics of the droplet is visualized in the same direction.
d
heights. Since, the effect of curvature of the surfaces is pronounced in The curvature effect is quantified for different curvature ratios δ = D0
S
the azimuthal direction, the maximum droplet spreading diameter which relates the initial droplet and curvature diameters. The diameters
(dmax ) on the curved surfaces is measured along that direction as shown of semi-cylindrical surfaces chosen for most of the current investiga-
in Fig. 3 using the correlation dmax = (Ds × θs )/2 . The residence time of tions are 8.002 mm and 11.986 mm such that δ is 0.29 and 0.19 respec-
droplets in the film boiling regime is obtained by calculating the time tively. Moreover, for high values of curvature ratios (δ = 0.58 & 1.16 ),
elapsed between impact and re-bounce. The uncertainty analysis of the the flat region adjacent to the curved surface interfere on the droplet
present experimental data is performed using Kline-McClintok method extension, thus proper quantification becomes difficult. The droplet
[45] and the maximum uncertainties obtained for d 0, vd, We , dmax and dispensing and visualization of the impact are performed as detailed in
TS are ± 2.06%, ± 3%, ± 4.76%, ± 2.06% and ± 0.68% respectively. the previous section.
In-order to understand the general behavior of the droplet impact,

255
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

Fig. 7. Superheated droplet dynamics on the convex surfaces for We = 38.63.

we performed experiments on a heated convex surface. The droplet experiments in the current study focused on temperatures exceeding the
impact on a convex surface with δ = 0.29 for different temperatures is static Leidenfrost temperature of TL = 170 °C [46]. Following Bernardin
recorded as shown in Fig. 4. The surface temperature is raised sys- et al. [6] the static Leidenfrost point is calculated by plotting the eva-
temically from 125 °C to 250 °C and the droplets are released from the poration time of a water droplet against the surface temperature on a
same height to maintain the impact velocity. Nucleate or contact flat test surface. Since, the droplet is impinged with certain finite ve-
boiling regimes are identified for surface temperatures, TS = 125 °C and locity the droplet rebounds from the surface in the film boiling regime.
150 °C (see Fig. 4(a) and (b)). Here, the vapor bubbles are generated at This is owing to the conversion of radial momentum into vertically
isolated nucleation sites and rise up to the top surface of the droplet and upward momentum and the pressure exerted by vapor layer underneath
the droplet boils off. In this regime, the drop is in contact with the on the drop as seen in Fig. 4(d) and (e) for TS = 225 °C and TS = 250 °C.
surface during the entire evaporation and a heating duration of 71 ms is Thus, rebounding is ensued for a shorter interval of 12 ms after the
required for nucleation at TS = 125 °C. As the surface temperature is collision due to the rapid formation of a cushioning layer owing to the
further raised to 175 °C, transitional boiling occurs and the vapor bub- high degree of superheat [5,12].
bles coalesce to form slugs and prevent the liquid surface contact in- Variation in droplet shape with respect to time, after impact on a
termittently due to which the droplet hovers above the surface as ex- hot convex surface (δ = 0.29 ) at TS = 275 °C and We = 5.41 in azimuthal
hibited in Fig. 4(c). A further increase in surface temperature led to film and axial directions are illustrated in Fig. 5(a)–(e) and (f)–(j) respec-
boiling, where the droplet levitates on the vapor layer and instead of tively. A droplet suspended in air exists in spherical shape prior to the
boiling it evaporates slowly as in Fig. 4. Note that, most of the collision as seen in Fig. 5(a) and (f). When the drop is in contact with

256
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

lamella is noticed as in Fig. 6. Resultantly, a δ = 0.29 surface facilitate


more deformation of droplet in the azimuthal direction. Eventually, the
inertial forces overpower the surface tension forces and droplet frag-
ments. In contrast, for δ = 0.19 surface, the deformation of the droplet
in the azimuthal direction is lower compared to that of a surface having
small diameter. Concurrently, the surface tension force is sufficient to
envelop the liquid mass as a single unit.
The variation in droplet behavior for two different surface tem-
peratures at a constant Weber number of 38.63 for convex surfaces with
δ = 0.19 and 0.29 is shown in Fig. 7. It is observed that, at TS = 175 °C
droplet disintegration occurred during the spreading phase irrespective
of the curvature ratio. Furthermore, the secondary droplets formed are
rebounded individually from the surface. Importantly, this phenom-
enon is not observed for a surface temperature of 250 °C where the
droplet bounced as a single mass. Conversion of water as vapor at
heterogeneous locations on the periphery during the drop spreading
phase coupled with dominant inertia forces rupture the droplets at
Fig. 8. Variation in β with K at different Weber numbers on the convex sur- TS = 175 °C. However, when the temperature is elevated to 250 °C rapid
faces. creation of vapor cushion enables the surface tension to suppress the
inertia force and the droplet mass bounced up as a single unit. Further,
the number of secondary drops generated are more on the surface with
δ = 0.19 owing to large spreading of the drop and concomitant eleva-
tion in the heat transfer area. Nonetheless, the formation of a stable
vapor layer at TS = 250 °C avoided contact boiling of the droplet. As a
result, the fragmentation of the liquid droplets ceases and it rebounded
vertically-upwards as in Fig. 7.
A drop after impact spreads radially into a disc shape when collided
over a flat surface [47]. However, the spreading on a curved surface is
asymmetrical due to the anisotropy in surface geometry [22,40]. In
addition, the maximum spreading diameter influences the available
heat transfer area during the droplet-surface collision period. Here, the
effect of surface superheat on droplet deformation is quantified by
plotting the maximum spread factor β = dmax / d 0 in the azimuthal di-
T −T
rection with a dimensionless superheat factor K = TS − satT
which is the
sat f
ratio of surface superheat to liquid sub-cooling. The laboratory and
saturation temperatures are taken as Tf = 28 °C and Tsat = 100 °C re-
spectively. The variation of β with different K values for impact Weber
Fig. 9. Effect of K on τr∗ for different Weber numbers on the convex surfaces. numbers 16.48 and 53.27 is depicted in Fig. 8. This describes the effect
of surface temperature on the maximum spread factor of the droplet.
the hot curved surface, the droplet spreads and yields the shape of The maximum spread factor β is almost invariant at different surface
curvature as in Fig. 5(b) and (g) at 4 ms. After a duration of 8 ms the temperatures except at higher values of K. This is owing to the dom-
droplet initiates to recoil as depicted in Fig. 5(c) and (h). The pressure inance of surface tension over the inertial forces due to rapid formation
exerted by the vapor layer as well as redirection of radial momentum of of the insulating vapor layer with an increase in surface superheat.
retracting fluid exert an effective force in the upward direction on the Further, the maximum spread factor is found to increase with a rise in
droplet, see Fig. 5(d) and (i). Eventually, the drop lifts-off from the Weber number as depicted in Fig. 8. Furthermore, the maximum spread
surface (Ref. Fig. 5(e) and (j)) owing to the momentum gained in the factor is simultaneously elevated with the curvature ratio since, the
upward vertical direction. It is evident that, spreading pattern of the preferential spreading is higher for a lower diameter surface. In addi-
drop is anisotropic as the curvature is not symmetric. tion, the gravity forces assist the extension of droplets in the azimuthal
For a detailed investigation, the droplet time evolution on convex direction for a semi-cylindrical surface.
surfaces with different impact conditions are imaged for a surface Next, the residence time of the colliding droplets on convex surfaces
temperature of 275 °C which is well above the static Leidenfrost tem- is quantified. The time duration up to which a droplet is in contact with
perature. To have a representative comparison of the droplet dynamics the heated surface is an important parameter with regard to heat
on convex surfaces, we chose curvature ratios of 0.29 and 0.19 and transfer process. During the film boiling of liquid droplets, the heat is
Weber numbers 28.73 and 53.27. The time evolution of the droplet transferred by conduction and radiation through the vapor layer which
deformation and disintegration are monitored and sequence of images is significantly influenced by the residence time, τr∗ [14,16,34]. It is
are depicted in Fig. 6. At We = 28.73, the spreading pattern of the noticed that, τr∗ is almost uniform irrespective of the surface superheat
droplet is almost similar for both the curved surfaces, however the re- as shown in Fig. 9. However, when the semi-cylindrical surface dia-
bound of the droplet from the surface is different. This is attributed to meter is more (δ = 0.19 ) the residence time is more as given in Fig. 9 for
reduction in recoiling of the drop on a δ = 0.29 surface which is having both Weber numbers. Since, the downward extension is more for
a high curvature due to which the impact kinetic energy overcomes the smaller diameter surface (δ = 0.29 ), the gravity force adversely affects
surface tensional forces. The spreading increases with rise in the Weber the droplet retraction. Subsequently, the rebounding of droplet occur
number to 53.27, due to high kinetic energy attained by the drop. after a longer delay period.
Nevertheless, the droplet disintegrated into secondary droplets on
δ = 0.29 surface during rebounding. Indeed, similar behavior is not
observed on a surface having δ = 0.19 although thinning of the liquid

257
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

Fig. 10. Morphology of water droplet on concave surfaces at temperature TS = 275 °C.

4. Droplet impact dynamics on superheated concave surfaces Next, the maximum spread factor of the droplet is estimated for
concave surfaces. The variation of β in the azimuthal direction against
In this section, the impingement dynamics of droplets on super- K on the concave surfaces for Weber numbers 16.48 and 53.27 is depicted
heated concave surfaces is described. Initially, two semi-circular cav- in Fig. 12. At superheated conditions, the vapor layer formation is so
ities of diameter 7.92 mm and 11.85 mm are considered for drop im- rapid, so that not much difference is noticed between bouncing pat-
pingement. Note that, the curvature ratios (δ = 0.29 and δ = 0.19) of terns. Similar to convex surfaces, the maximum spread factor increases
these surfaces are same that of convex surfaces elaborated in the pre- with a rise in the Weber number and the surface superheat has negli-
vious section. Similar to convex surfaces, droplets are subjected to gible effect on it. Further, the maximum spread factor increased when
spreading, recoiling and rebounding on concave surfaces as well. As the the curvature ratio is decreased. This is attributed to the reduction in
droplet hits the substrate the fluid spread along both axial and azi- anisotropic spreading as the diameter rises. Most importantly, the
muthal directions. However, the droplet behavior on concave surfaces gravity force adversely affects the extension of drop in a concave geo-
is quite different in comparison to a convex counterpart. Fig. 10 ex- metry as the diameter reduces.
hibits, the droplet dynamics for We = 28.73 and 53.27 on a surface with Fig. 13 illustrates the variation of residence time for various surface
temperature 275 °C. The temporal development of drop geometry is temperatures. It is observed from Fig. 13 that, τr∗ is greater for the cavity
provided in Fig. 10 for both curvature ratios and We. It is noticed that, with larger radius for same We. It is further observed that the surface
2 ms after the impact the droplet spreads completely in the azimuthal superheat is not significantly influencing the residence time. The
direction and formation of vapor layer commences. When, the time asymmetry in spreading in both azimuthal and longitudinal direction
reaches 4 ms the recoiling phase begins and the liquid mass agglom- caused the difference in bouncing time [22] for the levitating droplets.
erate to the point of impact owing to the favorable surface tension and A special phenomenon known as liquid-jet formation during the
gravity forces. Here, the effect of curvature on the extension in azi- drop impact on flat surface for We = 62.82 , is reported earlier [14]. This
muthal direction is not significant as in the case of a convex surface. effect is owing to the compression of central vapor bubble by the high
Influence of surface superheat on the water droplet dynamics on kinetic energy drops. Subsequently, the compressed gas burst the dro-
concave surfaces is shown in Fig. 11. It is noteworthy that, droplet plet and escape through the bubble cavity region along the central line
disintegration occurred at TS = 175 °C during the spreading phase si- as shown in Fig. 14. Similar behavior is exhibited by the concave sur-
milar to that of convex surfaces and this effect decreases as temperature face for a temperature of 200 °C. The vapor inside the compressed
elevates. Secondary droplet formation is visible in the regime due to the bubble protrudes through the flattened droplet on a concave surface.
breakage of liquid lamella owing to high gas pressure generated. Fur- Eventually, the liquid at the top surface splashing as a jet as depicted in
ther, the droplet fragmentation is disappeared for TS > TL due to the Fig. 14. The Worthington’s jet formation on concave surface takes place
stable film boiling. Concomitantly at TS = 250 °C, the droplet move at a lower We compared to a flat substrate. However, such phenomenon
vertically upwards as a single unit and bouncing off is observed. is not observed for convex surfaces in the range of Weber numbers and

258
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

Fig. 11. Time evolution of drops on superheated concave surfaces at We = 38.63.

Fig. 12. Variation in β with K at different Weber numbers on concave surfaces. Fig. 13. Effect of K on τr∗ for different Weber numbers on concave surfaces.

surface temperatures that have studied. This indicates that the geo- impact on the hot plate irrespective of the surface curvature as seen in
metry of surface plays vital role in the liquid jet formation. Fig. 15. However, the timing of occurrence of these events is different
and is to be discussed in detail later. Moreover, Fig. 16 depicts the
variation in maximum spread factor with Weber number for all three
5. Comparison of impact dynamics on flat and curved surfaces surfaces at a temperature of 290 °C. Thereafter, the correlations of Akao
et al. (β = 0.613We 0.39 ) [29], Hatta et al. (β = 0.093We 0.74 + 1) [28] and
This section discusses about the relative differences in droplet im- Liang et al. (β = 0.788We 0.306) [16] for superheated flat surfaces are
pact dynamics on surfaces with different curvatures under identical compared against the results of present study. It is noteworthy that, the
conditions. The variation in droplet morphology on flat, convex and maximum spread factor, β is least for a concave surface and highest for
concave surfaces for We = 16.48 at TS = 275 °C is illustrated in Fig. 15. a convex surface. This is owing to preferential spreading of the droplet
The droplet undergo spreading, retraction and bouncing during the

259
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

Fig. 14. Central liquid jet on the flat surface and concave surfaces at TS = 200 °C.

Fig. 15. Morphology of impacted water droplet on different surfaces for We = 16.48 and TS = 275 °C.

260
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

along the curvature. Hence, the drop on a concave surface is expected


to extend more in the axial direction. The results predicted by the
empirical correlations [16,28,29] are higher in comparison to the cur-
rent experimental values. These experiments were performed on po-
lished smooth surfaces which are hydrophilic in nature. In contrast, the
surfaces employed for the present study is hydrophobic with a contact
angle, θc = 125.06° ± 3. This hydrophobic nature of metal surface en-
abled a lesser spreading of droplets as the interfacial forces are domi-
nant. Similar anomalous behavior was also reported by different groups
while studying the Liedenfrost propulsion [48] as well as droplet
bouncing [49] on hydrophobic surfaces.
Time evolution of the spread diameter, ds of a water droplet for a
surface temperature of 290 °C at an impact Weber number 28.73 is
quantified in Fig. 17. The drop after the impact, extends gradually and
reaches the maximum spreading diameter. Thereafter, it decreases due
to retraction of the fluid towards the point of impact and reaches a
minimum value and then the drop bounces off vertically from the
Fig. 16. Variation in β with We on the flat, convex and concave surfaces at surface. The recoil and rebound of droplet are owing to the surface
TS = 290 °C and predicted results from previous literature [16,28,29]. tensional forces as well as the pressure exerted by the vapor layer
formed between the drop and the surface [47]. Note that, after a
duration of ≈ 3.5 ms the droplet on a flat surface commence receding, in
contrast the spreading on the convex surface is continued. The favor-
able effect of gravity helps the drop to spread in the azimuthal direction
for convex surfaces. Moreover, the extension and receding phase on the
concave surfaces is much quicker than its flat and convex counterparts,
since the gravity significantly influences the vertical movement of fluid
in the azimuthal direction. It is noteworthy that, the rebound occurred
on the flat surface around 12 ms while that is 9 ms and 8 ms for convex
and concave surfaces respectively as illustrated in Fig. 17. This is in-line
with the earlier prediction of Liu et al. [22] for a textured curved sur-
faces kept at laboratory conditions.
On a flat surface, the drop after impact spreads radially (circular) in
a symmetric fashion. Meantime, a curved surface facilitates an asym-
metry in drop spreading (elliptical) which eventually lead to a reduc-
tion in the residence time [22]. It is observed from Fig. 15 that, the
maximum spreading on a superheated flat surface (δ ≈ 0 ) is occurred at
3.5 ms after the impact, however which is 4 ms on a convex surface
(δ = 0.29 ) and 3.25 ms on a concave surface (δ = 0.29 ) respectively. The
Fig. 17. Comparison of the change in drop diameter with respect to time on a variation in spreading time for concave/convex surfaces from a flat
flat, convex and concave surfaces at TS = 290 °C and We = 28.73 .
solid surface suggest that there exists a preferential spreading in azi-
muthal direction similar to the earlier observations [22]. In this sym-
metry-breaking surface, the retraction force developed on the drop is
Fr ≈ πDmax σ (1 − cosθc ) [22] where Dmax represents the major axis of
ellipsoidal shaped drop in the oblate position. The gravitational force
Fg ≈ ρD03 g acting on the rim of fluid in the azimuthal direction, where g
is the acceleration due to gravity. These forces contribute an additive
effect on the drop for a convex surface, Fd ≈ Fr + Fg and residence time
is least. In contrast, for a convex surface the effective force is
Fd ≈ Fr − Fg , and the residence time is further reduced. Thus, the dro-
plet rebound takes place at different times after the impact. Here, the
maximum residence time is observed for the flat surface (12.75 ms)
which is followed by convex (10.75 ms) and concave (9.75 ms) surfaces.
A concave surface displays ≈ 24% reduction in contact time in com-
parison to an equivalent flat surface while that for a convex surface is
≈ 16%. Variation of dimensionless residence time with Weber number
of flat and curved surfaces at TS = 275 °C is depicted in Fig. 18. The
contact time is greater for the flat surface and least for the concave
surface. The experimental values are compared against the correlations
Fig. 18. Variation in residence time τr with We on flat, convex and concave proposed by Chen et al. (τr = 1.12We 0.5 ) [34] and Liang et al.
surfaces at TS = 275 °C. Empirical predictions from previous literature [34,16] (τr = 1.032We 0.494 ) [16] for flat substrates. Since, the surfaces deployed
is plotted along with the current results. Here, the time is non-dimensionalized for the present examination are hydrophobic in nature, a subsequent
with convective time scale, τr = tr vd/ d 0 [4,15,16,50].
drop in the residence time of the droplet is noticed.

in the azimuthal direction in conjunction with a favorable gravity force 6. Conclusions


for a convex surface. The spreading in the azimuthal direction is lower
for the concave surface, since the fluid has to flow against the gravity In this study, the impact dynamics of milli-metric water droplets on

261
R. Simhadri Rajesh et al. Experimental Thermal and Fluid Science 101 (2019) 251–262

superheated curved surfaces is experimentally investigated using high- [17] A. Worthington, On the forms assumed by drops of liquids falling vertically on a
speed imaging technique. The major objective is to examine the effect horizontal plate, Proc. Royal Soc. London 25 (1876) 261–272.
[18] J.T. Pearson, D. Maynes, B.W. Webb, Droplet impact dynamics for two liquids
of surface curvature on the droplet behavior when the drops hit on impinging on anisotropic superhydrophobic surfaces, Exp. Fluids 53 (2012)
asymmetrically curved cylindrical surfaces. Droplet spreading pattern, 603–618.
recoiling and rebound are observed to be different for the flat, convex [19] J.C. Bird, R. Dhiman, H.-M. Kwon, K.K. Varanasi, Reducing the contact time of a
bouncing drop, Nature 503 (2013) 385–388.
and concave surfaces. The maximum spread factor is found to be [20] A.I. Aria, M. Gharib, Physicochemical characteristics and droplet impact dynamics
highest for convex surfaces due to preferential spreading of liquid in the of superhydrophobic carbon nanotube arrays, Langmuir 30 (2014) 6780–6790.
azimuthal direction as well as due to a favorable gravity force. Further, [21] C.E. Clavijo, J. Crockett, D. Maynes, Effects of isotropic and anisotropic slip on
droplet impingement on a superhydrophobic surface, Phys. Fluids 27 (2015)
an increase in diameter of convex surface reduced the droplet extension 122104.
while a concave surface enhances the drop spreading. Moreover, a re- [22] Y. Liu, M. Andrew, J. Li, J.M. Yeomans, Z. Wang, Symmetry breaking in drop
duction of 24% and 16% in contact time is observed for a concave and a bouncing on curved surfaces, Nature Commun. 6 (2015).
[23] J. Jung, S. Jeong, H. Kim, Investigation of single-droplet/wall collision heat transfer
convex profiles respectively in comparison to a flat substrate. The effect
characteristics using infrared thermometry, Int. J. Heat Mass Trans. 92 (2016)
of surface superheat in the film boiling regime on the maximum spread 774–783.
factor and contact time is found to be negligible. These factors increase [24] T. Tran, H.J. Staat, A. Prosperetti, C. Sun, D. Lohse, Drop impact on superheated
with the Weber number owing to a rise in impact kinetic energy. The surfaces, Phys. Rev. Lett. 108 (2012) 036101.
[25] J.H. Moon, M. Cho, S.H. Lee, Dynamic wetting and heat transfer characteristics of a
droplet disintegration during the spreading phase is occurring at a liquid droplet impinging on heated textured surfaces, Int. J. Heat Mass Transf. 97
lower Weber number on the convex surfaces in comparison to a flat (2016) 308–317.
surface. The central liquid jet formation is observed on flat and concave [26] S. Chen, V. Bertola, Jumps, somersaults, and symmetry breaking in leidenfrost
drops, Phys. Rev. E 94 (2016) 021102.
surfaces at a surface temperature of 200 °C due to the compressed gas [27] L. Wachters, N. Westerling, The heat transfer from a hot wall to impinging water
zone at the impact point. The mechanisms behind spreading asymmetry drops in the spheroidal state, Chem. Eng. Sci. 21 (1966) 1047–1056.
on heated curved surfaces and liquid jet formation are further need to [28] N. Hatta, H. Fujimoto, H. Takuda, K. Kinoshita, O. Takahashi, Collision dynamics of
a water droplet impinging on a rigid surface above the leidenfrost temperature, ISIJ
be investigated in detail. Int. 35 (1995) 50–55.
[29] F. Akao, K. Araki, S. Mori, A. Moriyama, Deformation behaviours of a liquid droplet
Acknowledgments impinging onto hot metal surface, Trans. Iron Steel Inst. Jpn. 20 (1980) 737–743.
[30] S. Chandra, C. Avedisian, On the collision of a droplet with a solid surface,
Proceedings of the Royal Society of London A: Mathematical, Physical and
The authors gratefully acknowledge the financial support provided Engineering Sciences, vol. 432, The Royal Society, 1991, pp. 13–41.
by Kerala State Council for Science, Technology and Environment [31] A. Karl, A. Frohn, Experimental investigation of interaction processes between
droplets and hot walls, Phys. Fluids 12 (2000) 785–796.
(SARD-Scheme), Government of Kerala, India and Center for
[32] S.G. Kandlikar, M.E. Steinke, A. Singh, Effects of weber number and surface tem-
Engineering Research and Development (Best Researcher Grant), perature on the boiling and spreading characteristics of impinging water droplets,
Government of Kerala, India for this research work. The authors sin- Proc. NHTC 1 (2001).
cerely thank Dr. Joy Mithra, IISER-Trivandrum, Kerala for helping with [33] V. Bertola, An experimental study of bouncing leidenfrost drops: comparison be-
tween newtonian and viscoelastic liquids, Int. J. Heat Mass Transf. 52 (2009)
SEM images. 1786–1793.
[34] R.-H. Chen, S.-L. Chiu, T.-H. Lin, Resident time of a compound drop impinging on a
References hot surface, Appl. Therm. Eng. 27 (2007) 2079–2085.
[35] G. Cossali, M. Marengo, M. Santini, Secondary atomisation produced by single drop
vertical impacts onto heated surfaces, Exp. Thermal Fluid Sci. 29 (2005) 937–946.
[1] A. Yarin, Drop impact dynamics: splashing, spreading, receding, bouncing, Annu. [36] G. Cossali, M. Marengo, M. Santini, Thermally induced secondary drop atomisation
Rev. Fluid Mech. 38 (2006) 159–192. by single drop impact onto heated surfaces, Int. J. Heat Fluid Flow 29 (2008)
[2] A. Moreira, A. Moita, M. Panao, Advances and challenges in explaining fuel spray 167–177.
impingement: How much of single droplet impact research is useful? Prog. Ener. [37] E.-S.R. Negeed, M. Albeirutty, Y. Takata, Dynamic behavior of micrometric single
Comb. Sci. 36 (2010) 554–580. water droplets impacting onto heated surfaces with tio 2 hydrophilic coating, Int. J.
[3] M. Marengo, C. Antonini, I.V. Roisman, C. Tropea, Drop collisions with simple and Therm. Sci. 79 (2014) 1–17.
complex surfaces, Curr. Opin. Coll. Inter. Sci. 16 (2011) 292–302. [38] L. Hung, S. Yao, Experimental investigation of the impaction of water droplets on
[4] C. Josserand, S. Thoroddsen, Drop impact on a solid surface, Ann. Rev. Fluid Mech. cylindrical objects, Int. J. Multiphase Flow 25 (1999) 1545–1559.
48 (2016) 365–391. [39] M. Pasandideh-Fard, M. Bussmann, S. Chandra, Simulating droplet impact on a
[5] S. Mitra, M.J. Sathe, E. Doroodchi, R. Utikar, M.K. Shah, V. Pareek, J.B. Joshi, substrate of arbitrary shape, Atomizat. Sprays 11 (2001).
G.M. Evans, Droplet impact dynamics on a spherical particle, Chem. Eng. Sci. 100 [40] G. Liang, Y. Guo, Y. Yang, S. Guo, S. Shen, Special phenomena from a single liquid
(2013) 105–119. drop impact on wetted cylindrical surfaces, Exp. Thermal Fluid Sci. 51 (2013)
[6] J. Bernardin, I. Mudawar, The leidenfrost point: experimental study and assessment 18–27.
of existing models, J. Heat Transfer 121 (1999) 894–903. [41] Y. Hardalupas, A. Taylor, J. Wilkins, Experimental investigation of sub-millimetre
[7] H.K. Cho, K.Y. Choi, S. Cho, C.-H. Song, Experimental observation of the droplet droplet impingement on to spherical surfaces, Int. J. Heat Fluid Flow 20 (1999)
size change across a wet grid spacer in a 6× 6 rod bundle, Nucl. Eng. Des. 241 477–485.
(2011) 4649–4656. [42] C.K. Chow, D. Attinger, Visualization and measurements of microdroplet impact
[8] M. Visaria, I. Mudawar, Application of two-phase spray cooling for thermal man- dynamics on a curved substrate, ASME/JSME 2003 4th Joint Fluids Summer
agement of electronic devices, IEEE Trans. Compon. Packag. Technol. 32 (2009) Engineering Conference, American Society of Mechanical Engineers, 2003, pp.
784–793. 405–413.
[9] R.H. Stanglmaier, J. Li, R.D. Matthews, The effect of in-cylinder wall wetting lo- [43] S. Bakshi, I.V. Roisman, C. Tropea, Investigations on the impact of a drop onto a
cation on the HC emissions from SI engines, Technical Report, SAE Technical Paper, small spherical target, Phys. Fluids 19 (2007) 032102.
1999. [44] W.G. Bae, K.Y. Song, Y. Rahmawan, C.N. Chu, D. Kim, D.K. Chung, K.Y. Suh, One-
[10] G. Grant, J. Brenton, D. Drysdale, Fire suppression by water sprays, Progress Energy step process for superhydrophobic metallic surfaces by wire electrical discharge
Combust. Sci. 26 (2000) 79–130. machining, ACS Appl. Mat. Int. 4 (2012) 3685–3691.
[11] J.G. Leidenfrost, On the fixation of water in diverse fire, Int. J. Heat Mass Transf. 9 [45] S.J. Kline, F. McClintock, Describing uncertainties in single-sample experiments,
(1966) 1153–1166. Mech. Eng. 75 (1953) 3–8.
[12] D. Quéré, Leidenfrost dynamics, Ann. Rev. Fluid Mech. 45 (2013) 197–215. [46] R.S. Rajesh, Experimental study of droplet impact on superheated surfaces, Master’s
[13] L. Rayleigh, On the capillary phenomena of jets, in: Proc. R. Soc. London, vol. 29, thesis, 2016.
1879, pp. 71–97. [47] M. Rein, Drop-surface Interactions vol. 456, Springer, 2014.
[14] T. Tran, H.J. Staat, A. Susarrey-Arce, T.C. Foertsch, A. van Houselt, H.J. Gardeniers, [48] G. Dupeux, P. Bourrianne, Q. Magdelaine, C. Clanet, D. Quéré, Propulsion on a
A. Prosperetti, D. Lohse, C. Sun, Droplet impact on superheated micro-structured superhydrophobic ratchet, Sci. Rep. 4 (2014).
surfaces, Soft Matter 9 (2013) 3272–3282. [49] D.J. Lee, Y.S. Song, Anomalous water drop bouncing on a nanotextured surface by
[15] G. Liang, Y. Guo, S. Shen, Y. Yang, Crown behavior and bubble entrainment during the leidenfrost levitation, Appl. Phys. Lett. 108 (2016) 201604.
a drop impact on a liquid film, Theor. Comput. Fluid Dyn. 28 (2014) 159–170. [50] M. Rieber, A. Frohn, A numerical study on the mechanism of splashing, Int. J. Heat
[16] G. Liang, S. Shen, Y. Guo, J. Zhang, Boiling from liquid drops impact on a heated Fluid Flow 20 (1999) 455–461.
wall, Int. J. Heat Mass Transf. 100 (2016) 48–57.

262

Das könnte Ihnen auch gefallen