Stability Analysis of Printed Liquid Elbows - ACS Publications

Dec 14, 2016 - pressure differences, liquids flowed into the concave elbow corner and formed bulges easily after being printed. However, the bulge for...
0 downloads 0 Views 1MB Size
Subscriber access provided by GAZI UNIV

Article

Stability Analysis of Printed Liquid Elbows Kuan-Ming Huang, Tsung-Yen Tsou, Chia-Wei Chang, and Ying-Chih Liao Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.6b03906 • Publication Date (Web): 14 Dec 2016 Downloaded from http://pubs.acs.org on December 20, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Langmuir is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Impact of Viscous Droplets on Superamphiphobic Surfaces Binyu Zhao,† Xiang Wang,‡ Kai Zhang,§ Longquan Chen,*,† Xu Deng,*,ǁ †

State Key Laboratory of Traction Power, Applied Mechanics and Structure Safety Key Laboratory of Sichuan Province, School of Mechanics and Engineering, Southwest Jiaotong University, Chengdu 610031, China ‡ Department of Mechanical Engineering, Dongguan University of Technology, Dongguan 523808, China § Wood Technology and Wood Chemistry, Georg-August-Universität Göttingen, Büsgenweg 4, Göttingen D-37077, Germany ǁ Institute of Fundamental and Frontier Sciences, University of Electronic Science and Technology of China, Chengdu 610054, China

Abstract The impact of a liquid droplet on a solid surface is one of the most common phenomena in nature and frequently encountered in numerous technological processes. Despite the significant progress on understanding droplet impact phenomena in the past one century, the impact dynamics, especially those coupling effects between liquid property and surface wettability on the impact process, is still less understood. In this work, we experimentally investigated the impact of viscous droplets on superamphiphobic surfaces, with the viscosity of liquids ranging from 0.89 mPa s to

150 mPa s. We show that the increase of liquid viscosity will slow down the impact

process and cause bouncing droplets to rebound lower and fewer times. The critical impact velocity, above which droplets can rebound from the superamphiphobic surface, is found to linearly increase with the liquid viscosity. We also show that the maximum spreading factor increases with Weber number or Reynolds number but decreases with liquid viscosity. Scaling analyses based on energy conservation were carried out to explain these findings and they were found to be in good agreement with our experimental results.

Keywords: droplet impact, viscosity, viscous dissipation, critical velocity, maximum spreading factor

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 23

1. Introduction The impact of liquid droplets on solid surfaces is of prominent importance as it is involved in numerous technological processes1, such as ink-jet printing2, spray painting and spin coating3, pesticides deposition4, and also recently in bloodstain pattern analysis5. Driven by potential applications in these fields aforementioned, droplet impact has been extensively investigated for more than one century since the pioneering work of Worthington in 1876

1, 6, 7

. Various impact phenomena, including

deposition, complete rebound, partial rebound and splashing, have been identified1, 8, 9. In particular, the rebound of droplets from solid surfaces has attracted considerable attention of surface engineers due to its significant role in self-cleaning10,

11

,

anti-icing12-14, anti-fogging11, 15 and water harvesting16, 17. It has been demonstrated that surface superhydrophobicity18,

19

, Leidenfrost

effect20, 21, substrate sublimation22, and ambient pressure reduction23 can facilitate droplet rebound. Among them, surface superhydrophobicity is the one that has been widely studied. Superhydrophobic surfaces are low energy surfaces decorated with micro-, nano- or micro-/nanostructures, which can entrap a thin layer of air beneath the droplet and cause it to rebound after impact24-28. The impact process of a droplet on a superhydrophobic surface can be described as follows. Upon contact with the surface, the inertial force forces the droplet to expand up to a maximum extent. In this spreading process, most of the kinetic energy is converted into the interfacial energy, which triggers the subsequent droplet retraction. It is noted that both the spreading and retraction processes are associated with viscous energy dissipation within the droplet29-34. Therefore, a droplet can rebound only when its kinetic energy is high enough to compensate the energy dissipated during impact9, 35-37. For low viscosity liquid, e.g. water, a critical impact velocity for droplet rebound on superhydrophobic surfaces can be obtained by balancing the kinetic energy with the stored interfacial energy

35

. In industrial applications, however, highly viscous liquids, such as silicon

oils and paints, are commonly encountered. To the best of our knowledge, the impact of highly viscous droplets on superhydrophobic or superamphiphobic surfaces has received little attention33.

ACS Paragon Plus Environment

Page 3 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

The maximum spreading diameter (  ) is an important parameter for the

control of droplet deposition4, 38. Normalizing it by the initial droplet diameter (  ) yields a dimensionless parameter - the maximum spreading factor -   =

 /  . A number of theoretical models have been proposed to explain its relationship to the Weber number ( =   /, where ,  ,  and  are the

density, radius, velocity and surface tension of the droplet, respectively) or Reynolds number (  =   / with  being the viscosity) in the literature7,

39-41

. If

capillary force dominates the dynamics, a balance between the kinetic energy and interfacial energy results in   ∝  /

predicts   ∝  /

41, 42

39

, while momentum conservation

. On the other hand, if viscous force dominates the

dynamics, a scaling of   ∝  /

39, 43, 44

or   ∝ /

45

can be obtained

by balancing the kinetic energy and viscous dissipation energy. Recently, Laan et al.29 rescaled the spreading factor using   ∝  ⁄ g ( #/ ), based on which

 Lee et al.46 further developed a scaling of (  −  )⁄ ∝  / g  (), where

g is a function of  #/ , g  is a function of  and  is the equilibrium spreading factor at zero impact velocity. With these two models, they collapsed experimental data obtained from the impact of low viscosity droplets ( ≤ 51 mPa s)

on different partial wetting surfaces (the equilibrium contact angle '() is in the range of 23° − 110°). However, if these scaling analyses can be applied to predict  

of the impact of viscous droplets on superamphiphobic surfaces is still an open question. In this work, we thus carried out an experimental investigation on the impact of

various viscous droplets on superamphiphobic surfaces. The viscosity of the liquids ranges from 0.89 mPa s to 150 mPa s and the impact velocity is varied from 0.05 m/s to 2.73 m/s, which correspond to  of 0.03 − 131 and  of 1.31 − 2275. We found that the impact process is slowed down by viscous dissipation and the critical

velocity for droplet rebound linearly increases with the liquid viscosity. The maximum spreading factor of various viscous droplets on superamphiphobic surfaces cannot be described by the scaling model of Laan et al.29 and Lee et al.46. Preliminary models based on energy conservation were proposed to explain these findings.

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2. Experimental Section 2.1 Surface and Liquids Superamphiphobic surfaces were fabricated on the silicon substrate by soot-templated structure comprising a fractal-like network with hydrophobized silica shell. Detailed fabrication information is described elsewhere47. Scanning electron microscopy (SEM, Quanta 200, FEI) images reveal that these nanoparticles form a fractal-like network (Fig. 1a), which is a desirable structure for superamphiphobicity 48

. Droplet impact experiments were carried out with pure water and various

glycerol-water mixtures. These liquids have slightly different surface tensions (63.0 – 71.8 mN/m) but much different viscosities (0.89 – 150 mPa s), which are obtained from the literatures49-51 and shown in Table 1. The equilibrium ('() ), advancing (' ) and receding ('. ) contact angles of these liquids on the superamphiphobic surface

were measured by the sessile drop technique using a commercial goniometer (DSA4,

Krüss GmbH, Germany). A 5 µL droplet was carefully deposited on the surface. The static (equilibrium) contact angle was obtained by fitting the drop profile using Laplace-Young method. By inflating or deflating the drop, we determined the advancing and receding contact angles when the contact line started to move. Each measurement was repeated at least at three different places on the surface. The error

bar is defined by the standard deviation. As shown in Table 1, '() is larger than 150° for all liquids (Fig. 1b) and the contact angle hysteresis (' − '. ) is always below 10°.

Fig. 1 The superamphiphobic surface. (a) SEM images of the superamphiphobic surface coated on silicon substrate. The brace indicates protrusions and the spacing B in between. (b) Contact angles of 5 µL droplets of pure water (Top) and 60 wt% glycerol-water mixture (Bottom) on the

ACS Paragon Plus Environment

Page 4 of 23

Page 5 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

superamphiphobic surface shown in (a).

2.2 Droplet Impact Experiments & Data Analysis The superamphiphobic surface was placed on a horizontal stage, above which

5 − 8 μL droplets (  ≈ 2 mm) were generated by pumping liquids through a steel

needle using a syringe pump. Needles with an outer diameter of ~0.24 mm were applied throughout the experiments unless otherwise specified. The impact velocity was varied from 0.05 m/s to 2.73 m/s by changing the releasing height. The corresponding Weber number is 0.03 − 131 and the Reynolds number is 1.31 −

2275. We recorded the impact process with a high-speed camera (Photron, Fastcam

Mini UX100) at 12,800 fps. Each impact experiment was repeated at least three times. The

recorded

videos

of

droplet

impact

were

processed

using

a

custom-programmed MATLAB (Math_Works Inc., USA) algorithm. To quantify the impact process, we evaluated the variation of contact angle ('), contact diameter ( ) and centroid height (2) of the impinging droplets (Fig. 2 & Fig. 3).

Table 1. The viscosity (3), density (4) and surface tension (5) of various glycerol−water mixtures and the corresponding advancing (67 ), receding (68) and equilibrium (69: ) contact angles of these liquids on the superamphiphobic surfaces at 25°C. 3 (mPa s)49

4 (kg/m3)50

5 (mN/m)51

67 (°)

68 (°)

69: (°)

30

2.2

1071

66.5

163±1

153±3

158±4

60

8.8

1151

64.6

162±1

152±2

157±4

80

46

1205

63.3

163±1

152±2

157±5

85

82

1219

63.1

160±1

153±2

156±5

87

150

1227

63.0

159±1

152±1

156±4

Fraction (wt%) 0

0.89

997

71.8

163±1

3. Results and Discussion 3.1 Effects of liquid viscosity on the impact process

ACS Paragon Plus Environment

159±1

161±1

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

When a liquid droplet impacts a non-wetting surface, it generally spreads on the surface and then retracts back. Eventually, the droplet can stick to or rebound from the surface depending on the impact velocity, surface wettability and liquid viscosity1, 9, 33, 36, 52

. We conducted a systematic study of droplet impact of pure water and various

glycerol-water mixtures on the superamphiphobic surface. We found that droplets could only rebound from the surface above a critical impact velocity, below which droplet deposition was observed. However, further increase of the impact velocity led to the occurrences of partial rebound, sticking and splashing of droplets. Fig. 2 shows the snapshots of droplets of water (left), 60 wt% (middle) and 85 wt% glycerol-water mixtures (right) impacting on the superamphiphobic surface at  = 0.24 8.8 mPa s . Fig. 5a & b compares the impact process of three liquids at

 = 0.89 m/s. Similar to bouncing droplets, a droplet with higher viscosity spreads

and retracts slower than droplets with low viscosities, and a smaller contact diameter is attained. It is also seen that the dynamic contact angle of all droplets is smaller than

150° during the whole impact process (Fig. 5b), indicating the penetration of liquids

into surface structure. Similar to other studies9, 33, droplets splashing was observed at very high velocity in our experiment. It happens at  ≳ 2.2 m/s for water and 2.5

m/s for 30 wt% glycerol-water mixture ( = 2.2 mPa s), while no splashing was observed for other viscous liquids in the range of impact velocity we studied. 2

b 160

0.89 mPa s 8.8 mPa s 82 mPa s

120

θ (°)

a Dc/D0 (mm)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 23

1

80 0.89 mPa s 8.8 mPa s 82 mPa s

40

0

0

2

4

6

8

10

t (ms)

12

14

16

0

0

2

4

6

8

10

t (ms)

12

14

16

Fig. 5 Time evolution of HI /HC (a) and 6 (b) for the impact of droplets with different

ACS Paragon Plus Environment

Page 11 of 23

viscosities on the superamphiphobic surfaces at BC = C. ab F/G.

3.2 The critical velocity for droplet rebound

A droplet can rebound from a solid surface when its kinetic energy is high enough to compensate energy dissipated during impact. Therefore, there should exist a critical

impact velocity ? , above which droplet rebound is possible to occur. In this study,

this critical velocity is much smaller than the threshold velocity for liquid penetration into surface structures (0.40-0.69 m/s, depending on liquid properties) and splashing

( ≥2.2 m/s, depending liquid viscosity) during impact. Fig. 6 summarizes ? for

various viscous droplets on the superamphiphobic surface. It is seen that ? linearly increases with .

0.3

0.2

Vc (m/s)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

0.1

0.0

0

40

80

120

160

µ (mPa s)

Fig. 6 Plot of the critical impact velocity BI for droplet rebound as a function of the liquid

viscosity 3. The dashed line is the best linear fitting.

For

the

impact

of

low

viscosity

droplets

on

superhydrophobic

or

superamphiphobic surfaces, viscous dissipation inside the droplet is always negligible as the Reynolds number is much larger than 1

58

. The kinetic energy of impinging

droplet scales as [  and the surface energy can be stored during impact is estimated as  |cos' − cos '. |

9, 35, 36

. The initial kinetic energy should be higher

than the stored surface energy for the droplet to rebound. Thus, one obtains a critical velocity for droplet rebound ? = e|cos' − cos '. |/ . For water, the above

scaling analysis predicts ? = 0.029 m/s on our superamphiphobic surface, which is consistent with the value (~0.031 m/s) that we measured in the experiment (see Fig.

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 23

6). However, for liquids with higher viscosities, the initial kinetic energy is mainly dissipated by viscosity, so the above model will underestimate the critical impact velocity. Herein, we derive a critical impact velocity for viscous droplets based on energy

conservation. We assume that an impinging droplet with ? takes off from the

superamphiphobic surface with a much lower velocity, i.e. the kinetic energy of

rebounding droplet is negligible. Moreover, at low impact velocity a droplet slightly deforms during impact and always takes a spherical shape after impact 9. Therefore, we can also assume that droplet’s surface energy is the same before and after impact. Furthermore, although the flow field inside impinging droplet during spreading and retraction is different, the characteristic properties of the flow, e.g. the thickness of the boundary layer, the duration, should be in the same order of magnitude. As a result, the kinetic energy of an impinging droplet is dissipated by liquid viscosity. The viscous dissipation energy fg can be estimated as59

fg = h o hm i jk jA ≈ ikl? n

in which i is the viscous dissipation function i = p

qgr

qs

+

qgs

u

qgr

qr qs

≈  v Ox L

w

(1) 

(2)

where k is the characteristic volume for viscous dissipation, l? is the contact time,

and δ is a characteristic length in the vertical direction.

Pasandideh-Fard et al.45 demonstrated that viscous dissipation mainly occurs in

the boundary layer with a thickness of y = 2  ⁄√2 and thus the characteristic  volume can be estimated as k = z  y. They scaled the time needed for an  

impinging droplet to reach the maximum spreading as A  ≈ 8  ⁄3 , and thus one could estimate the contact time as l? ≈ 2A  . However, we found this scaling

strongly overestimates the contact time, e.g. it predicts a value of ~44 ms at an impact velocity of 0.24 m/s while the actual l? is 9.5 – 11.5 ms as shown in Fig. 3. In contrast, we found that the contact time can be well estimated by l? ≈  ⁄. Thus,

the viscous dissipation energy can be further expressed as

ACS Paragon Plus Environment

Page 13 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

 fg =   √2 { |

(3)

To promote a droplet rebound from the superamphiphobic surface, the kinetic energy,

f} = z [  /12, should be higher than the viscous dissipation energy, i.e. f} > fg ,

which leads to

 > ? =

ƒ ~€‚

? ≈



„RO



(4)

where the maximum spreading factor   =  /  . At low impact velocity,   is close to 1 (see Fig. 7), and thus we obtain ~

„RO

(5)

This linear relationship between … and  matches our experimental results in

Fig. 6. A linear regression of the data in Fig. 6 provides a slope of 1.24 m/(Pa s2), which is also close to the value in Eq. (5).

3.3 The maximum spreading factor

The maximum spreading factor,   , which is defined as the ratio of the

maximum spreading diameter (  ) to the initial droplet diameter before impact (  ), is the key parameter controlling droplet dynamics in practical applications4, 5, 60

and much effort has been devoted to understand the maximum spreading during droplet impact29, 32, 38-40, 42, 43, 45, 46. A number of theoretical models, which depend on the type of dissipation energy resisting impact, have been proposed to describe the relationship between   and  or Re

38, 39, 43, 45

. However, these models are

only applicable for specific liquids or surfaces. Recently, Laan et al.29 and Lee et al.46

developed a universal rescaling of droplet impact respectively, with which they collapsed   of different viscous droplets ( ≤ 51 mPa s) on hydrophilic and

hydrophobic surfaces (23° ≤ '() ≤ 110°) into one master curve. Here, we will check

if these scaling analyses are also applicable for impact of various viscous droplets (

is up to 150 mPa s) on the superamphiphobic surfaces ('() = 156°~161°).

Fig. 7a & b plot   of liquids with various viscosities as a function of 

and  , respectively.   increases with  and  but decreases with the ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 23

increase of liquid viscosity, which is similar to the results obtained on hydrophobic and hydrophilic surfaces39,

. We also found that we cannot collapse   of

43

droplets with different viscosity on superamphiphobic surfaces by either the scaling model of Laan et al.29 or Lee et al.46 (see Fig. S1 & Fig. S2 in the Supporting Information). This is plausible as a simple scaling analysis is hard to capture all energy involved in droplet impact. To understand the impact dynamics of viscous liquids on superamphiphobic surfaces, we balance all energy terms involved before impact and at the maximum spreading. Prior to impact, the kinetic energy is f} = z [  /12, the surface

energy of the droplet is f† = z   and the interfacial energy of the surface in air is  f‡ ≈ z  ˆ‰Š /4, where ˆ is the solid fraction of the nanostructure and ‰Š is

the solid/air interfacial tension. At the maximum spreading, we treated the droplet as a

circular cylindrical disc with a thickness of ℎ. Thus, the surface energy of the droplet  scales as f†P ≈ Œ(2 − ˆ )z  /4 + z  ℎ and the interfacial energy of the

 surface in liquid is f‡P ≈ z  ˆ‰Ž /4, where ‰Ž is the solid/liquid interfacial

tension. With energy conservation equation f} + f† + f‡ = f†P + f‡P + fg , we obtain [

(

(

 √(

[ + 1 − cos 'U )  − ( + 6)  + 4 = 0

(6)

with cos 'U = ˆ(cos'‘ + 1) − 1, where 'U is the contact angle of a droplet stays on

the structured surface in the Cassie state53, and '‘ is the Young’s contact angle of

the liquid on the smooth surface. On our superamphiphobic surfaces, static droplets always stay in the Cassie state and thus 'U ≈ '() . Solving the above equation, we

find

  = ’ •– cosœ [ “#…˜™ š 

in which œ = arccos ž18’√N—–  [

•–

((“”)

√N—–

“#…˜™ š–›

((“”)Ÿ

–›

  .

ACS Paragon Plus Environment

(7)

Page 15 of 23

a

6 5

βmax

3

1/2

2

1 -2 10

b

We

0.89 mPa s 2.2 mPa s 8.8 mPa s 46 mPa s 82 mPa s 150 mPa s

4

10

6 5

-1

10

0

10

We

3

1

10

2

1/4

Re

0.89 mPa s 2.2 mPa s 8.8 mPa s 46 mPa s 82 mPa s 150 mPa s

4

βmax

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

2

1 10

0

10

1

10

Re

2

10

3

Fig. 7 The maximum spreading factors as a function of We (a) and Re (b). The dashed lines are determined from Eq. (7). The solid lines in (a) and (b) are depicted for scaling behaviors of

  =  ⁄ and   =  ⁄ , respectively.

As shown in Fig. 7a & b, Eq. (7) agrees our experimental data both in trend and order of magnitude. The agreement is the best for highly viscous droplets at high impact velocity, while for low viscosity droplets or low impact velocity, Eq. (7)

overestimates   . This phenomenon may be due to two reasons. First, for low viscosity droplets, a capillary wave can be generated after contact with the solid surface at 1 ≲  ≲ 10 9. This capillary wave propagates along the droplet and

leads to the formation of an air cavity at the droplet center when the droplet reaches the maximum spreading. In our model, additional surface energy and viscous dissipation due to this strong deformation were not considered, which would result in

a larger   . Second, at the maximum spreading an internal circular flow exists near the contact line 39. This internal kinetic energy was also not considered in our model

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 23

and thus would cause an overestimation of   .

Eq. (7) can be further simplified to the scaling in the literature. If  ≪

√2(1 − cos '() ), the viscous dissipation within the droplet can be neglected during impact. At high impact velocity with  ≫ 6, we obtain   = £ cosœ ∝  / [ #…˜™ š 

(

–›

(8)

On the other hand, if  ≫ 6 and  ≫ √2(1 − cos '() ) , the viscous

dissipation dominates the dynamics and Eq. (7) reduces to    = e2√2 ∝  / [

(9)

For comparison, the above scaling is also plotted in Fig. 7 (the solid lines). We found that the maximum spreading factor of all liquids we used cannot be described

by   ∝ /. In contrast, the scaling of   ∝  / can capture the trend

of   for highly viscous liquids. These findings highlight the importance to consider viscous dissipation in the course of droplet impact dynamics.

Several theoretical models for the maximum spreading factor could also be found in the literature

30, 38, 45, 59

. However, we found that none of these models could

provide a good estimation for   on the superamphiphobic surface (see Fig. S3 S5 in the Supporting Information).

Pasandideh-Fard et al.45 suggested that the viscous dissipation in a spreading droplet mainly occurs within a boundary layer. They neglected part of the surface energy of the droplet at the maximum spreading and calculated the viscous dissipation with a much longer time scale as we discussed before. As shown in Fig. S3 in the

Supporting Information, their model provides a good estimation of   for low viscosity liquids at high , but it underestimates   for highly viscous liquids.

The maximum deviation between the theoretical prediction and our experimental

result is up to 42% (see liquid with  = 150 mPa s in Fig. S3). Li et al.30 assumed that the viscous dissipation occurs within the entire nanodroplet and estimated the

spreading time as the time to change the droplet height from  to h. However, such

assumption is not appropriate for the impact of millimeter-sized droplet in our

ACS Paragon Plus Environment

Page 17 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

experiment, and thus their model is not applicable for our study. Indeed, a maximum deviation of 55% is found between their model and our experimental data (see Fig. S4 in the Supporting Information). Ukiwe et al.38 modified the theoretical model of Pasandideh-Fard et al.45 by considering the complete surface energy of the impinging

droplet. This model overestimates   for low viscosity liquids with a maximum

deviation of 28% at high , but it underestimates   for highly viscous liquids with a maximum deviation of 41% (see Fig. S5 in the Supporting Information).

In contrast, we refined the model of Pasandideh-Fard et al.45 by using an appropriate characteristic time for droplet spreading and a complete estimation of the surface energy at the maximum spreading. As a result, a good agreement between the model and experimental result for   , especially for high viscosity droplets, is

reached. The maximum deviation between the theoretical value and the experimental result is 25% for water, and it decreases to 11% for liquid with viscosity of 150 mPa s, as shown in Fig. 7.

4. Conclusion In conclusion, we comparatively investigated the impact dynamics of various viscous droplets on superamphiphobic surfaces. We found that droplets with high viscosity spread and retract slower, take off the surface later and eventually rebound lower and less times than droplets with low viscosity. This phenomenon demonstrates that the viscous dissipation within droplet can strongly affect the impact process. Experiments also showed that the critical impact velocity for droplet rebound linearly increases with the liquid viscosity, which can be explained by a scaling argument. Furthermore, we found that the maximum spreading factor of various viscous droplets on superamphiphobic surfaces cannot be described by the scaling analyses recently reported in the literatures29, 46. Based on energy conservation, we developed a simple model to explain our experimental data. Since the impact of liquid droplets on solid surfaces is commonly encountered in industry, our results can shed light on the use of superamphiphobic surfaces in practical applications.

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60



Page 18 of 23

Author information

Corresponding Author *E-mail: [email protected] (L.Q.C.); [email protected] (X.D.) Notes The authors declare no competing financial interest.



Acknowledgements

This research was supported by the National Young 1000 Talents Plan, the Young 1000 Talents Plan of Sichuan province, Sichuan Province Science Foundation for Youths (Grant No. 2016JQ0050) and the University Initiative Grant No. A0920502051607-1 of SWJTU. X.D acknowledges the National Natural Science Foundation of China (Grant No.21603026).



Associated Content

Supporting information is available free of charge

via the Internet at

http://pubs.acs.org/.

References 1. Yarin, A. L., Drop impact dynamics: Splashing, spreading, receding, bouncing... Annual Review of Fluid Mechanics 2006, 38, 159-192. 2. Galliker, P.; Schneider, J.; Eghlidi, H.; Kress, S.; Sandoghdar, V.; Poulikakos, D., Direct printing of nanostructures by electrostatic autofocussing of ink nanodroplets. Nat Commun 2012, 3, 890. 3. Rioboo, R.; Tropea, C.; Marengo, M., Outcomes from a drop impact on solid surfaces. Atomization and Sprays 2001, 11, (2), 155-165. 4. Bergeron, V.; Bonn, D.; Martin, J. Y.; Vovelle, L., Controlling droplet deposition with polymer additives. Nature 2000, 405, (6788), 772-775. 5. Laan, N.; de Bruin, K. G.; Slenter, D.; Wilhelm, J.; Jermy, M.; Bonn, D., Bloodstain Pattern Analysis: implementation of a fluid dynamic model for position determination of victims. Scientific Reports 2015, 5, 11461. 6. Worthington, A. M., On the Forms Assumed by Drops of Liquids Falling Vertically on a Horizontal Plate. Proceedings of the Royal Society of London 1876, 25, (171-178), 261-272. 7. Josserand, C.; Thoroddsen, S. T., Drop Impact on a Solid Surface. Annual Review of Fluid Mechanics 2016, 48, (1), 365-391. 8. Martin, R., Phenomena of liquid drop impact on solid and liquid surfaces. Fluid

ACS Paragon Plus Environment

Page 19 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Dynamics Research 1993, 12, (2), 61-93. 9. Chen, L. Q.; Xiao, Z. Y.; Chan, P. C. H.; Lee, Y. K.; Li, Z. G., A comparative study of droplet impact dynamics on a dual-scaled superhydrophobic surface and lotus leaf. Applied Surface Science 2011, 257, (21), 8857-8863. 10. Lu, Y.; Sathasivam, S.; Song, J.; Crick, C. R.; Carmalt, C. J.; Parkin, I. P., Robust self-cleaning surfaces that function when exposed to either air or oil. Science 2015, 347, (6226), 1132-1135. 11. Wisdom, K. M.; Watson, J. A.; Qu, X.; Liu, F.; Watson, G. S.; Chen, C. H., Self-cleaning of superhydrophobic surfaces by self-propelled jumping condensate. Proc Natl Acad Sci U S A 2013, 110, (20), 7992-7. 12. Mishchenko, L.; Hatton, B.; Bahadur, V.; Taylor, J. A.; Krupenkin, T.; Aizenberg, J., Design of ice-free nanostructured surfaces based on repulsion of impacting water droplets. ACS Nano 2010, 4, (12), 7699-707. 13. Maitra, T.; Tiwari, M. K.; Antonini, C.; Schoch, P.; Jung, S.; Eberle, P.; Poulikakos, D., On the nanoengineering of superhydrophobic and impalement resistant surface textures below the freezing temperature. Nano Lett 2014, 14, (1), 172-82. 14. Zhang, Q.; He, M.; Chen, J.; Wang, J.; Song, Y.; Jiang, L., Anti-icing surfaces based on enhanced self-propelled jumping of condensed water microdroplets. Chemical Communications 2013, 49, (40), 4516-4518. 15. Boreyko, J. B.; Chen, C. H., Self-propelled dropwise condensate on superhydrophobic surfaces. Phys Rev Lett 2009, 103, (18), 184501. 16. Chen, X.; Wu, J.; Ma, R.; Hua, M.; Koratkar, N.; Yao, S.; Wang, Z., Nanograssed Micropyramidal Architectures for Continuous Dropwise Condensation. Advanced Functional Materials 2011, 21, (24), 4617-4623. 17. Ju, J.; Bai, H.; Zheng, Y.; Zhao, T.; Fang, R.; Jiang, L., A multi-structural and multi-functional integrated fog collection system in cactus. Nat Commun 2012, 3, 1247. 18. Gauthier, A.; Symon, S.; Clanet, C.; Quere, D., Water impacting on superhydrophobic macrotextures. Nat Commun 2015, 6, 8001. 19. Liu, Y. H.; Moevius, L.; Xu, X. P.; Qian, T. Z.; Yeomans, J. M.; Wang, Z. K., Pancake bouncing on superhydrophobic surfaces. Nature Physics 2014, 10, (7), 515-519. 20. Zhang, W.; Yu, T.; Fan, J.; Sun, W.; Cao, Z., Droplet impact behavior on heated micro-patterned surfaces. Journal of Applied Physics 2016, 119, (11), 114901. 21. Lee, D. J.; Song, Y. S., Anomalous water drop bouncing on a nanotextured surface by the Leidenfrost levitation. Applied Physics Letters 2016, 108, (20), 201604. 22. Antonini, C.; Bernagozzi, I.; Jung, S.; Poulikakos, D.; Marengo, M., Water drops dancing on ice: how sublimation leads to drop rebound. Phys Rev Lett 2013, 111, (1), 014501. 23. Schutzius, T. M.; Jung, S.; Maitra, T.; Graeber, G.; Köhme, M.; Poulikakos, D., Spontaneous droplet trampolining on rigid superhydrophobic surfaces. Nature 2015, 527, (7576), 82-85. 24. Liu, Y.; Whyman, G.; Bormashenko, E.; Hao, C.; Wang, Z., Controlling drop

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

bouncing using surfaces with gradient features. Applied Physics Letters 2015, 107, (5), 051604. 25. Vollmer, D.; Butt, H.-J., Materials science: Droplets leap into action. Nature 2015, 527, (7576), 41-42. 26. Chen, L. Q.; Wu, J.; Li, Z. G.; Yao, S. H., Evolution of entrapped air under bouncing droplets on viscoelastic surfaces. Colloids and Surfaces a-Physicochemical and Engineering Aspects 2011, 384, (1-3), 726-732. 27. de Ruiter, J.; Lagraauw, R.; van den Ende, D.; Mugele, F., Wettability-independent bouncing on flat surfaces mediated by thin air films. Nat Phys 2015, 11, (1), 48-53. 28. Bouwhuis, W.; van der Veen, R. C. A.; Tran, T.; Keij, D. L.; Winkels, K. G.; Peters, I. R.; van der Meer, D.; Sun, C.; Snoeijer, J. H.; Lohse, D., Maximal Air Bubble Entrainment at Liquid-Drop Impact. Physical Review Letters 2012, 109, (26), 264501. 29. Laan, N.; de Bruin, K. G.; Bartolo, D.; Josserand, C.; Bonn, D., Maximum Diameter of Impacting Liquid Droplets. Physical Review Applied 2014, 2, (4), 044018. 30. Li, X.-H.; Zhang, X.-X.; Chen, M., Estimation of viscous dissipation in nanodroplet impact and spreading. Physics of Fluids 2015, 27, (5), 052007. 31. Hao, C. L.; Li, J.; Liu, Y.; Zhou, X. F.; Liu, Y. H.; Liu, R.; Che, L. F.; Zhou, W. Z.; Sun, D.; Li, L.; Xu, L.; Wang, Z. K., Superhydrophobic-like tunable droplet bouncing on slippery liquid interfaces. Nature Communications 2015, 6, 7986. 32. Lee, J. B.; Derome, D.; Dolatabadi, A.; Carmeliet, J., Energy Budget of Liquid Drop Impact at Maximum Spreading: Numerical Simulations and Experiments. Langmuir 2016, 32, (5), 1279-1288. 33. Deng, X.; Schellenberger, F.; Papadopoulos, P.; Vollmer, D.; Butt, H. J., Liquid Drops Impacting Superamphiphobic Coatings. Langmuir 2013, 29, (25), 7847-7856. 34. Bartolo, D.; Josserand, C.; Bonn, D., Retraction dynamics of aqueous drops upon impact on non-wetting surfaces. Journal of Fluid Mechanics 2005, 545, 329-338. 35. Reyssat, M.; Pepin, A.; Marty, F.; Chen, Y.; Quere, D., Bouncing transitions on microtextured materials. Europhysics Letters 2006, 74, (2), 306-312. 36. Bartolo, D.; Bouamrirene, F.; Verneuil, E.; Buguin, A.; Silberzan, P.; Moulinet, S., Bouncing or sticky droplets: Impalement transitions on superhydrophobic micropatterned surfaces. Europhysics Letters 2006, 74, (2), 299-305. 37. Hao, P. F.; Lv, C. J.; Niu, F. L.; Yu, Y., Water droplet impact on superhydrophobic surfaces with microstructures and hierarchical roughness. Science China-Physics Mechanics & Astronomy 2014, 57, (7), 1376-1381. 38. Ukiwe, C.; Kwok, D. Y., On the Maximum Spreading Diameter of Impacting Droplets on Well-Prepared Solid Surfaces. Langmuir 2005, 21, (2), 666-673. 39. Clanet, C.; Béguin, C.; Richard, D.; Quéré, D., Maximal deformation of an impacting drop. Journal of Fluid Mechanics 2004, 517, 199-208. 40. Vaikuntanathan, V.; Sivakumar, D., Maximum Spreading of Liquid Drops Impacting on Groove-Textured Surfaces: Effect of Surface Texture. Langmuir 2016, 32, (10), 2399-2409.

ACS Paragon Plus Environment

Page 20 of 23

Page 21 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

41. Eggers, J.; Fontelos, M. A.; Josserand, C.; Zaleski, S., Drop dynamics after impact on a solid wall: Theory and simulations. Physics of Fluids 2010, 22, (6), 062101. 42. Bennett, T.; Poulikakos, D., Splat-quench solidification: estimating the maximum spreading of a droplet impacting a solid surface. Journal of Materials Science 1993, 28, (4), 963-970. 43. Lee, J. B.; Derome, D.; Guyer, R.; Carmeliet, J., Modeling the Maximum Spreading of Liquid Droplets Impacting Wetting and Nonwetting Surfaces. Langmuir 2016, 32, (5), 1299-1308. 44. Fedorchenko, A. I.; Wang, A.-B.; Wang, Y.-H., Effect of capillary and viscous forces on spreading of a liquid drop impinging on a solid surface. Physics of Fluids 2005, 17, (9), 093104. 45. Pasandideh ‐ Fard, M.; Qiao, Y. M.; Chandra, S.; Mostaghimi, J., Capillary effects during droplet impact on a solid surface. Physics of Fluids 1996, 8, (3), 650-659. 46. Lee, J. B.; Laan, N.; de Bruin, K. G.; Skantzaris, G.; Shahidzadeh, N.; Derome, D.; Carmeliet, J.; Bonn, D., Universal rescaling of drop impact on smooth and rough surfaces. Journal of Fluid Mechanics 2015, 786, R4. 47. Deng, X.; Mammen, L.; Butt, H. J.; Vollmer, D., Candle Soot as a Template for a Transparent Robust Superamphiphobic Coating. Science 2012, 335, (6064), 67-70. 48. Butt, H.-J.; Semprebon, C.; Papadopoulos, P.; Vollmer, D.; Brinkmann, M.; Ciccotti, M., Design principles for superamphiphobic surfaces. Soft Matter 2013, 9, (2), 418-428. 49. Sheely, M. L., Glycerol Viscosity Tables. Industrial & Engineering Chemistry 1932, 24, (9), 1060-1064. 50. Bosart, L. W.; Snoddy, A. O., Specific Gravity of Glycerol. Industrial & Engineering Chemistry 1928, 20, (12), 1377-1379. 51. Romero, C. M.; Paez, M. S., Surface tension of aqueous solutions of alcohol and polyols at 298.15 K. Physics and Chemistry of Liquids 2006, 44, (1), 61-65. 52. Rioboo, R.; Voue, M.; Vaillant, A.; De Coninck, J., Drop impact on porous superhydrophobic polymer surfaces. Langmuir 2008, 24, (24), 14074-14077. 53. Cassie, A. B. D.; Baxter, S., Wettability of porous surfaces. Transactions of the Faraday Society 1944, 40, (0), 546-551. 54. Zorba, V.; Stratakis, E.; Barberoglou, M.; Spanakis, E.; Tzanetakis, P.; Anastasiadis, S. H.; Fotakis, C., Biomimetic Artificial Surfaces Quantitatively Reproduce the Water Repellency of a Lotus Leaf. Advanced Materials 2008, 20, (21), 4049-4054. 55. Deng, T.; Varanasi, K. K.; Hsu, M.; Bhate, N.; Keimel, C.; Stein, J.; Blohm, M., Nonwetting of impinging droplets on textured surfaces. Applied Physics Letters 2009, 94, (13), 133109. 56. Chen, L.; Li, Z., Bouncing droplets on nonsuperhydrophobic surfaces. Physical Review E 2010, 82, (1), 016308. 57. Roman, B.; Hannes, A.; Samir, C.; Stefan, C.; Franz, K.; Bernhard, J., Resonant pressure wave setup for simultaneous sensing of longitudinal viscosity and sound

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

velocity of liquids. Measurement Science and Technology 2013, 24, (12), 125101. 58. Richard, D.; Quéré, D., Bouncing water drops. EPL (Europhysics Letters) 2000, 50, (6), 769-775. 59. Chandra, S.; Avedisian, C. T., On the Collision of a Droplet with a Solid Surface. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 1991, 432, (1884), 13-41. 60. Sun, J.; Bao, B.; He, M.; Zhou, H.; Song, Y., Recent Advances in Controlling the Depositing Morphologies of Inkjet Droplets. ACS Applied Materials & Interfaces 2015, 7, (51), 28086-28099.

ACS Paragon Plus Environment

Page 22 of 23

Page 23 of 23

TOC

Vc (m/s)

0.3 0.2

Vc≈(9/4ρD0)/µ

0.1 0.0 0

βmax

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

40

6 5 4 3

100

101

120

160

Re1/4

0.89 mPa s 2.2 mPa s 8.8 mPa s 46 mPa s 82 mPa s 150 mPa s

2

1

80

µ (mPa s)

102

Re

103

TOC. The critical velocity for droplet rebound linearly increases with droplet viscosity, and the maximum spreading factor increases with Reynolds number.

ACS Paragon Plus Environment