You are on page 1of 9

J_ID: APPLAB DOI: 10.1063/5.

0045281 Date: 8-April-21 Stage: Page: 7 Total Pages: 8

PROOF COPY [APL21-AR-00653]

Applied Physics Letters ARTICLE scitation.org/journal/apl

AUTHOR QUERY FORM

Please provide your responses and any corrections


Journal: Phys. Fluids
by annotating this PDF and uploading it to AIP’s

eProof website as detailed in the Welcome email.

Article Number: POF22-AR-02572

Dear Author,
Below are the queries associated with your article; please answer all of these queries before sending the proof back to AIP.

Article checklist: In order to ensure greater accuracy, please check the following and make all necessary corrections before returning your
proof.
1. Is the title of your article accurate and spelled correctly?
2. Please check affiliations including spelling, completeness, and correct linking to authors.
3. Did you remember to include acknowledgment of funding, if required, and is it accurate?

Location in Query / Remark: click on the Q link to navigate


article to the appropriate spot in the proof. There, insert your comments as a PDF annotation.

AQ1 Please check that the author names are in the proper order and spelled correctly. Also, please ensure that each
author’s given and surnames have been correctly identified (given names are highlighted in red and surnames
appear in blue).
AQ2 Please check and confirm the presentation for Refs. 10 and 34.
AQ3 Please provide the page number for Ref. 11 and also provide DOI for the same..
AQ4 Please provide the publisher name for Ref. 26.
AQ5 We were unable to locate a digital object identifier (doi) for Ref(s). 28 and 29. Please verify and correct author names
and journal details (journal title, volume number, page number, and year) as needed and provide the doi. If a doi is
not available, no other information is needed from you. For additional information on doi’s, please select this link:
http://www.doi.org/.

Please confirm ORCIDs are accurate. If you wish to add an ORCID for any author that does not have one, you may do so
now. For more information on ORCID, see https://orcid.org/.

G. Gillot-
J.-M. Genevaux-
L. Simon - 0000-0001-6115-9160
L. Benyahia - 0000-0002-8910-4097

Please check and confirm the Funder(s) and Grant Reference Number(s) provided with your submission:

Please add any additional funding sources not stated above:

Thank you for your assistance.

ID: aipepub3b2server Time: 18:05 I Path: D:/AIP/Support/XML_Signal_Tmp/AI-APL#210696


J_ID:
J_ID:APPLAB
PHFLE6 DOI:
DOI:10.1063/5.0045281
10.1063/5.0098642 Date:
Date:8-April-21
2-July-22 Stage:
Stage: Page:
Page: 71 Total
Total Pages:
Pages: 89

PROOF
PROOF COPY
COPY [APL21-AR-00653]
[POF22-AR-02572]

Applied Physics
Physics of FluidsLetters ARTICLE
ARTICLE scitation.org/journal/apl
scitation.org/journal/phf

1 Fast dynamics of surfactant probed


2 by the acoustics of a drop impact
4 Cite as: Phys. Fluids 34, 000000 (2022); doi: 10.1063/5.0098642
7
5 Submitted: 11 May 2022 . Accepted: 23 June 2022 .
6 Published Online: 0 Month 0000
8
AQ1 G. Gillot,1,2 J.-M. Ge
 nevaux,2,a) L. Simon,2 and L. Benyahia1,a)
9
10 AFFILIATIONS
1
11  cules et Mate
Institut des Mole  riaux du Mans (IMMM), UMR 6283 CNRS–Universite  du Mans. 1, Avenue Olivier Messiaen,
12 72085 Le Mans cedex 9, France
2
13 Laboratoire d’Acoustique de l’Universite du Mans (LAUM), UMR CNRS 6613, Institut d’Acoustique- Graduate School (IA-GS),
14 72085 Le Mans cedex 9, France

15 a)
Authors to whom correspondence should be addressed: Jean-Michel.Genevaux@univ-lemans.fr
16 and Lazhar.Benyahia@univ-lemans.fr

ABSTRACT
17 Adding a surfactant to water leads to changes in the outcome of a water drop impacting on the solution such as the dynamics of the Rayleigh
18 jet, and the same is true for the bubbles entrainment. The resulting acoustic signal is, therefore, modified in the presence of a surfactant and
19 is found to be related to the fast dynamics features of the latter. To this end, the airborne acoustic signal is synchronized with hydrodynamic
20 images, recorded by a high-speed camera, of a water drop impacting aqueous solutions with varying concentrations of three different surfac-
21 tants. It is found that the starting time of the acoustic events shows a maximum around the third of the critical micellar concentration inde-
22 pendently of the surfactant chemistry. This feature is related to the variation of the Rayleigh jet maximum height resulting mainly from a
23 subtle balance between an increase in viscosity and concentration in addition to Marangoni flows that affect the acceleration of the Rayleigh
24 jet and modify its initial speed.
Published under an exclusive license by AIP Publishing. https://doi.org/10.1063/5.0098642

25 INTRODUCTION As the sound produced by a drop impact depends on the post- 45

26 A drop impact on a liquid surface leads to a succession of impact dynamics, it is straightforward to deduce that changing the 46
27 hydrodynamic events such as the formation of a cavity, which liquid properties leads to modifications of the emitted sound. A well- 47
28 retracts to project a liquid jet, called the Rayleigh or Worthington known example concerns the “regular entrainment” regime for which 48
29 jet.1 During these events, an air bubble can be entrapped in the liq- addition of surfactant prevents the entrapment of a bubble and, there- 49
30 uid, generating a typical acoustic signal.2 The time and the way the fore, mutes the impact.16 Concerning the higher energy impact regime 50
31 bubbles are entrapped in the liquid depend on the diameter and the studied in Ref. 6, the addition of a surfactant leads to new mechanisms 51
32 velocity of the drop. Different bubble entrapment regimes are of bubble entrainment with different acoustic signatures.17 52
33 highlighted by Pumphrey et al.3 Among them, the so-called Surfactants play a major role when they are involved in sub- 53
34 “irregular entrainment” regime corresponds to the high energy second hydrodynamic events.18–20 Their fast adsorption and the 54
35 impact and consists of bubble entrainment by secondary droplets induced Marangoni flows can affect the behavior of the interface dur- 55
36 during the Rayleigh jet collapse.2,4,5 For this regime, a particular con- ing these events. Studying these changes can help better understanding 56
37 figuration, reported in Ref. 6, leads to bubble entrainment by a the surfactant fast dynamics and even developing new techniques to 57
38 highlighted new mechanism, called “liquid zip-like flow.” measure the dynamic surface tension.21,22 58
39 With non-pure water in the pool, the hydrodynamic events may Indeed, the amphiphilic character of surfactants confers them a 59
40 vary due to changes in liquid properties, such as surface tension or vis- specific ability to partition on the surface and in the bulk to decrease 60
41 cosity.7–11 Likewise, the addition of polymers or fibers to the aqueous the free energy by decreasing the surface tension while maximizing the 61
42 solution also modifies the jet dynamics,12 so does the use of non- entropy of the system. Above a critical concentration, called critical 62
43 Newtonian fluids.13,14 Similar changes are observed underwater for micellar concentration (CMC), micelles form in the bulk as the surfac- 63
44 more complex fluids.15 tant density at the interface reaches its maximum coverage and 64

Phys. Fluids 34, 000000 (2022); doi: 10.1063/5.0098642 34, 000000-1


Published under an exclusive license by AIP Publishing

ID: aipepub3b2server Time:


ID: aipepub3b2server Time: 18:05
12:22 II Path:
Path: D:/AIP/Support/XML_Signal_Tmp/AI-APL#210696
D:/AIP/Support/XML_Signal_Tmp/AI-PHF#221109
J_ID:
J_ID:APPLAB
PHFLE6 DOI:
DOI:10.1063/5.0045281
10.1063/5.0098642 Date:
Date:8-April-21
2-July-22 Stage:
Stage: Page:
Page: 72 Total
Total Pages:
Pages: 89

PROOF
PROOF COPY
COPY [APL21-AR-00653]
[POF22-AR-02572]

Applied Physics
Physics of FluidsLetters ARTICLE
ARTICLE scitation.org/journal/apl
scitation.org/journal/phf

65 consequently a plateau in surface tension.23 From the dynamics of sur- Br€uel and Kjaer (Type 2690). The microphone is positioned 68 mm 121
66 factant organization arise several properties that impact any flow of from the point of the impact. The signals are recorded with a 122
67 the fluid. First, the viscosity of the solution changes more noticeably Picoscope 4262 digital oscilloscope with a sampling rate of 500 kHz. 123
68 on passing the CMC as the hydrodynamic volume of a micelle is A video of the interface is taken with a Photron FastCam SAX2 124
69 much larger than the sum of the individual surfactants. Second, the fast camera, which is equipped with a Sigma Macro 105 mm lens. A 125
70 dynamics of absorption and reorganization of the surfactants at the backlight is provided by a 50 W LED RS Pro lamp. The sampling fre- 126
71 interface generate Marangoni flows, which inevitably affect the cover- quency is 20 kHz. 127
72 age or spread of the surface compared to its surfactant-free homolog.24 The acoustic and video acquisitions are triggered by an infrared 128
73 This paper presents a detailed study of the effect of adding surfac- sensor module XCSource TE174, fixed just underneath the hose and 129
74 tants on the acoustic signal as well as the related hydrodynamic events connected to an Arduino Mega card. The origin of time is set to be the 130
75 with a special focus on the Rayleigh jet features. Three surfactants with moment of the initial impact of the falling drop on the liquid surface. 131
76 different dynamics at the water/air interface are selected. The results Each signal is denoised with a spectral subtraction algorithm to 132
77 highlight a universal effect of surfactants on the studied phenomena reduce the background noise.25,26 Then, different acoustic events are 133
78 that is correlated with the CMC and Marangoni effect induced by the detected by a home-made Python algorithm, which uses a sliding win- 134
79 flow of the surfactant at the moving interface. dow to calculate the local energy along the signal and then detects the 135
local maxima above a given threshold. These maxima correspond to 136
80 MATERIAL AND METHODS
the initial impact and the beginning of different acoustic events 137
81 A Harvard Apparatus Ph.D. Ultra syringe pump is used to gener- afterward. 138
82 ate a distilled water drop with a flow rate of 10 ml/h. It is equipped
83 with a 60 ml syringe connected to an Interchim Teflon cylindrical hose RESULTS 139

84 with internal and external diameters of 1 and 1.5 mm, respectively. In the “irregular entrainment” regime, the airborne acoustic sig- 140
85 The drop falls into a Plexiglas tank of an aqueous solution with exter- nal of a drop impact on a liquid surface shows two main parts: the ini- 141
86 nal dimensions of 160  160  110 mm3 with 10 mm thick walls. The tial impact and the bubble sound event.6 The initial impact part shows 142
87 tank is filled until its maximum height and lies on blocks of solid poly- only a sharp pressure impulse due to the drop impact on the liquid 143
88 mer foam to reduce vibrations transmitted from the exterior to the surface without any specific frequency. The bubble sound part shows 144
89 tank. The drop diameter is 3.70 6 0.05 mm. The drop is released at one or several damped sines, characteristic of free vibration of a bub- 145
90 74 cm from the liquid surface, so the speed of the impact, measured on ble. These bubbles are entrained in the liquid during the Rayleigh jet 146
91 the video footage, is 3.32 6 0.05 m s1. collapse, mostly because of the impact of the secondary droplets 147
92 The complete setup is fixed to a metallic structure situated in a detached from the jet. 148
93 semi-anechoic room with a cutoff frequency of 100 Hz. The tempera- Fifty drop impact sounds and one video are recorded for each 149
94 ture is quite constant at 19  C. concentration. An overlay of airborne signals for different concentra- 150
95 The tank is filled with different solutions of a given surfactant in tions of SDS is shown in Fig. 1. As the initial impact part shows no dif- 151
96 distilled water at various concentrations. Three surfactants used are as ferences between signals, the bubble sound part shows a greater 152
97 follows: (1) an anionic surfactant: sodium dodecyl sulfate (SDS) pur- variability with the SDS concentration. However, the acoustic events 153
98 chased from Fisher Scientific as an aqueous solution at 20 wt. % (Lot: are gathered in several distinct groups for a given concentration bath 154
99 142978), (2) a nonionic surfactant: polyethylene glycol dodecyl ether (see Fig. S1 in the supplementary material). For example, Fig. 1(c) 155
100 (Brij-35) purchased from Aldrich (Lot: S08035-503), and (3) a cationic shows three distinct signal clusters, around 0.15, 0.17, and 0.20 s. 156
101 surfactant: cetyltrimethylammonium bromide (CTAB) purchased Indeed, for the same concentration, the bubble entrainment, leading to 157
102 from Acros Organics (Lot: A011462701). They all are used as received the acoustic events, occurs globally at the same time for all measure- 158
103 without any further purification. ments, but with slight variations due to little differences in the second- 159
104 The concentration is adjusted to cover a range from 0.02 to 12 ary droplets detachment. 160
105 times the surfactant CMC. After each concentration, the bath is left at Indeed, as reported in Ref. 17, the initial impact part of the signal 161
106 least 20 min at rest before the measurements. does not show significant variation when a surfactant is added to 162
107 For each concentration, a sample of the bath is collected. Its equi- water. However, the bubble sound part displays important changes in 163
108 librium surface tension is measured with the bubble profile method the presence of a surfactant as the number and the starting time of the 164
109 using an automatic drop tensiometer (TRACKER, from Teclis acoustic events evolve with the surfactant concentration. Thus, in Fig. 165
110 Scientific, France). In addition, the dynamic surface tension is mea- 1, the number and starting time of the acoustic events evolve from 166
111 sured with the bubble pressure method (PocketDyne BP2100 from only one event, around 0.125 s for a concentration of 0.02  CMC, to 167
112 Kr€uss, Germany). The characteristic time of measurement is fixed at three events for higher concentrations, with an increasing starting 168
113 90 ms with this last method. time, and then back to only one event, around 0.16 s, for a concentra- 169
114 For a given surfactant concentration, the acoustic signal in air is tion of 1.18  CMC. 170
115 recorded for 50 impacts. At least 20 s separate two successive impacts As the starting time of the acoustic events forms distinct groups, 171
116 to ensure the rest of the interface. A supplementary material measure- they can be easily identified with a simple clustering algorithm. 172
117 ment is realized by combining acoustic and video recordings. Different acoustic events for each concentration are, thus, defined by a 173
118 The acoustic signal of the impact is recorded in the air with a mean starting time with a standard deviation, indicating the low vari- 174
119 GRAS microphone (Type 40 BP 1/4 in). It is connected to a GRAS ability between measurements. Figure 2 shows the starting times of the 175
120 preamplifier (Type 26AC 1/4 in), linked to a conditioning amplifier acoustic events as a function of the concentration for three different 176

Phys. Fluids 34, 000000 (2022); doi: 10.1063/5.0098642 34, 000000-2


Published under an exclusive license by AIP Publishing

ID: aipepub3b2server Time:


ID: aipepub3b2server Time: 18:05
12:22 II Path:
Path: D:/AIP/Support/XML_Signal_Tmp/AI-APL#210696
D:/AIP/Support/XML_Signal_Tmp/AI-PHF#221109
J_ID:
J_ID:APPLAB
PHFLE6 DOI:
DOI:10.1063/5.0045281
10.1063/5.0098642 Date:
Date:8-April-21
2-July-22 Stage:
Stage: Page:
Page: 73 Total
Total Pages:
Pages: 89

PROOF
PROOF COPY
COPY [APL21-AR-00653]
[POF22-AR-02572]

Applied Physics
Physics of FluidsLetters ARTICLE
ARTICLE scitation.org/journal/apl
scitation.org/journal/phf

FIG. 1. Overlay of 50 microphone signals


of a drop impact on aqueous solutions
of SDS at a concentration of (a) 0.02
 CMC, (b) 0.12  CMC, (c) 0.24  CMC,
(d) 0.59  CMC, and (e) 1.18  CMC.
Each impact is represented in a different
color.

177 surfactants: SDS, CTAB, and Brij-35. As mentioned before, three dif- concentrations. This behavior is similar for the three surfactants. 210
178 ferent surfactants display various acoustic events in the concentrations Moreover, it is noticeable that the jets obtained with Brij-35 are glob- 211
179 range. Even though the reproducibility of the hydrodynamic events ally higher than those obtained with SDS and CTAB for the same 212
180 and the related acoustic signatures is unfortunately not favorable CMC ratio. To our knowledge, no similar observation has been 213
181 enough, a common behavior is observable in the three graphs. Indeed, reported in the literature. 214
182 the starting times of events increase globally and then decrease with To go further in the analysis of the hydrodynamics of the 215
183 the concentration, with a critical point at about CMC/3, independently Rayleigh jet, we have also determined its speed and acceleration at its 216
184 of the chemistry of the surfactant. This identification is supported by emergence. 217
185 other investigations such as the observation of the Rayleigh jet pre- The jet initial speed is measured in several frames when it 218
186 sented hereafter [Figs. 3 (Multimedia view) and 4]. appears above the surface on the video footage. The results are shown 219
187 Different observations may highlight this transition concentra- in Fig. 5. Again, the behavior is similar for the three surfactants. The 220
188 tion: (i) following a particular event, the first one in this case (䊉 sym- initial speed is quite constant below the CMC, with little variations 221
189 bols in Fig. 2); (ii) the time amplitude of the acoustic events that are mostly due to the jet tip dynamics. Above the CMC, the speed 222
190 indicated by the external envelope of the data. Both approaches show decreases with the concentration until stabilization for high concentra- 223
191 a transition concentration around CMC/3, even though for SDS, tions that is slightly higher for Brij-35. The jet speed is about the same 224
192 method I gives a slight height transiting concentration. Furthermore, for SDS and CTAB. 225
193 the concentration transition can be seen in the raw data of Fig. 2, espe- The acceleration, which is considered positive in the downward 226
194 cially when observing the events after CMC. Indeed, the starting time direction, is measured during the jet rise. As shown in Ref. 27, the jet 227
195 of events clearly decreases with the concentration for the three surfac- acceleration is high during the beginning of the jet rise, then decreases 228
196 tants, while the time spreading of the events reduces. rapidly to stabilize for the rest of the rise. Similar behavior is observed 229
197 Different acoustic events are generated during the Rayleigh jet for our measurements, so we report only the steady value of the accel- 230
198 collapse so depend on its dynamics.6,17 Five different families were eration. The results, higher than 9.81 m s2, are plotted in Fig. 6. For 231
199 identified in Ref. 17 and described regarding their hydrodynamic fea- the three surfactants, the acceleration decreases with the concentration 232
200 tures and acoustic signature in both the air and underwater. The video and then increases around the CMC. However, the minimum value is 233
201 footage, see “(Multimedia view),” shows indeed different Rayleigh jet quite different from one surfactant to another. The highest value is 234
202 dynamics, especially its maximum height, depending on the surfactant obtained with SDS and the lowest one with Brij-35. 235
203 concentration, as shown in Fig. 3 for SDS. It is noticeable that the
204 height variation is not monotonous with the concentration. DISCUSSION 236

205 The maximum height of the Rayleigh jet, observed on the video The particular behavior of the starting time of the acoustic events 237
206 associated with each concentration, is plotted vs the concentration for is, undoubtedly, linked to the variation of the maximum height of the 238
207 the three surfactants in Fig. 4. Similarly to the starting time of the Rayleigh jet. Indeed, the acoustic events are generated by the impact of 239
208 acoustic events, the maximum height increases with the concentration the secondary droplets on the cavity resulting from the disappearance 240
209 until a maximum and then decreases toward a plateau for higher of the jet.6 The first secondary droplet generally detaches when the jet 241

Phys. Fluids 34, 000000 (2022); doi: 10.1063/5.0098642 34, 000000-3


Published under an exclusive license by AIP Publishing

ID: aipepub3b2server Time:


ID: aipepub3b2server Time: 18:05
12:22 II Path:
Path: D:/AIP/Support/XML_Signal_Tmp/AI-APL#210696
D:/AIP/Support/XML_Signal_Tmp/AI-PHF#221109
J_ID:
J_ID:APPLAB
PHFLE6 DOI:
DOI:10.1063/5.0045281
10.1063/5.0098642 Date:
Date:8-April-21
2-July-22 Stage:
Stage: Page:
Page: 74 Total
Total Pages:
Pages: 89

PROOF
PROOF COPY
COPY [APL21-AR-00653]
[POF22-AR-02572]

Applied Physics
Physics of FluidsLetters ARTICLE
ARTICLE scitation.org/journal/apl
scitation.org/journal/phf

is at its maximum height. The higher the jet, the higher the position of 242
the detachment of the secondary droplets, the later the impacts of 243
these secondary droplets, and the later the acoustic events. The mea- 244
surements indeed show a later starting time for the highest Rayleigh 245
jets. 246
If we ignore the viscothermal losses during its development, the 247
maximum height of the Rayleigh jet mostly depends on its initial 248
velocity and on the forces acting on it (corresponding to its accelera- 249
tion).28 However, as seen in Figs. 5 and 6, both the initial speed and 250
the acceleration of the jet change according to the fraction of CMC. 251
Moreover, neither the initial speed nor the acceleration follows the 252
same trend as the maximum height. Indeed, as the jet maximum 253
height shows an increase and a decrease with a maximum at about 254
one third of the CMC, the initial velocity is quite constant until 255
approximately the CMC, then decreases sharply. Concerning the 256
acceleration, although it shows a decrease and an increase for the three 257
surfactants, the position of the minimum does not fit completely with 258
the position of the maximum hmax . Thus, neither the initial velocity 259
nor the acceleration can explain separately the behavior of hmax . 260
However, by considering the jet simply as an object projected ver- 261
tically upward with an initial velocity Vi and undergoing an accelera- 262
tion a, the evolution of its height hjet as a function of the time t follows 263
a parabolic curve, expressed as 264

a t2
hjet ¼  þ Vi t þ h0 ; (1)
2
with h0 being the height at t ¼ 0, corresponding also to the time at 265
which the initial velocity Vi is measured. In our configuration, h0 is 266
about 8 mm. Thus, the maximum height hmax is reached for a time 267
t ¼ Vai , which gives 268

Vi 2
hmax ¼ þ h0 : (2)
2a
2
Figure 7 shows hmax as a function of V2ai . Equation (2) is represented by 269
a black line and shows quite a satisfying agreement with the experi- 270
mental data with a slight deviation, which can be explained by the fol- 271
lowing considerations. The points grayed out at the right of the graph, 272
corresponding to heights lower than predicted by the model, are due 273
to cases when a secondary droplet detaches from the jet during its 274
riseT. This leads to energy losses and, therefore, to lower maximum 275
height considering the initial speed and acceleration. 276
It may also appear that the behavior of the SDS deviates slightly 277
more from the prediction of Eq. (2). In fact, with the exception of the 278
two blue points on the far right, the remaining points are reasonably 279
clustered around the theoretical prediction, given the measurement 280
noise. 281
Given the consistency of the model of Eq. (2) with the experi- 282
mental data, it can be concluded that the maximum height of the jet is 283
driven by a combination between its initial velocity and the accelera- 284
tion it undergoes. 285
As the behavior of hmax is explained by the jet initial velocity and 286
acceleration, the behavior of these two parameters can be explained by 287
considering the liquid properties, therefore linking them to hmax , and 288

FIG. 2. Starting time of the microphone acoustic events as a function of the CMC by extension to the acoustic event starting times. 289
fraction for three surfactants: SDS, CTAB, and Brij-35. 䊉, , 3, and x symbols First, the initial velocity displays a common behavior for the three 290
are related to the first, second, third, and fourth acoustic events, respectively. surfactants, i.e., no great changes before the CMC, and then a 291

Phys. Fluids 34, 000000 (2022); doi: 10.1063/5.0098642 34, 000000-4


Published under an exclusive license by AIP Publishing

ID: aipepub3b2server Time:


ID: aipepub3b2server Time: 18:05
12:22 II Path:
Path: D:/AIP/Support/XML_Signal_Tmp/AI-APL#210696
D:/AIP/Support/XML_Signal_Tmp/AI-PHF#221109
J_ID:
J_ID:APPLAB
PHFLE6 DOI:
DOI:10.1063/5.0045281
10.1063/5.0098642 Date:
Date:8-April-21
2-July-22 Stage:
Stage: Page:
Page: 75 Total
Total Pages:
Pages: 89

PROOF
PROOF COPY
COPY [APL21-AR-00653]
[POF22-AR-02572]

Applied Physics
Physics of FluidsLetters ARTICLE
ARTICLE scitation.org/journal/apl
scitation.org/journal/phf

FIG. 3. Footage of the Rayleigh jet at its


maximum height for water with SDS at
concentrations of (a) 0.02  CMC at
t ¼ 98 ms, (b) 0.12  CMC at t ¼ 83 ms,
(c) 0.24  CMC at t ¼ 98 ms, (d) 0.59
 CMC at t ¼ 111 ms, and (e) 1.18
 CMC at t ¼ 106 ms. Multimedia view:
https://doi.org/10.1063/5.0098642.1

292 significant decrease with the concentration from the CMC. Since the before the CMC. Thus, the surface tension forces cannot explain the 312
293 flow velocity is directly involved in the viscous dissipation of the fluid, particular behavior of the acceleration. Moreover, studies report that 313
294 it is tempting to correlate its behavior with that of the viscosity of sur- the equilibrium surface tension has no significant influence on the 314
295 factant solutions. Indeed, the viscosity is quite constant under the sur- maximum height of the Rayleigh jet 12 induced by an impact of a solid 315
296 factant CMC and then rises significantly above the CMC.29,30 This is projectile on a non-Newtonian fluid. 316
297 due to the formation of micelles in the liquid, which increases the In order to verify the weak influence of surface tension forces on 317
298 hydrodynamic radius of the suspended particle. As shown by Ref. 28, the maximum height of the Rayleigh jet, similar measurements are 318
299 viscothermal losses are great during the jet ejection. These losses can performed for water drop impact on water/ethanol mixtures, see Fig. 319
300 be increased by the increase in viscosity resulting from the increasing S2 in the supplementary material. No specific trend of hmax is observed 320
301 surfactant concentration, leading to lower kinetic energy for the jet as a function of surface tension, which confirms that hmax is influenced 321
302 and, therefore, a lower initial speed. by other surfactant features. 322
303 Concerning the acceleration, its value depends on all the forces During the cavity and jet formations, a new surface is created. As 323
304 acting on the jet. The forces, which can be identified, are gravity and the surfactant adsorption at the interface is not instantaneous, this 324
305 surface tension forces. As the gravity forces do not change with the new surface is necessarily poorer in the surfactant. As a result, the sur- 325
306 concentration, because the variations of the solution density are quite face tension is inevitably higher than the surface around and at rest. A 326
307 negligible, a change of surface-tension forces is considered to explain gradient of surface tension is then created at the air/water interface, 327
308 the changes in the acceleration. However, equilibrium surface tension which leads to the apparition of tangential flows along the interface, 328
309 is known to decrease with the surfactant concentration until the CMC, called Marangoni flows. These flows are directed from the point of 329
310 where it becomes constant for increasing concentrations. The accelera- lowest to the point of highest surface tension. In the case of the 330
311 tion displays instead of a decrease than an increase, with the minimum Rayleigh jet, its surface is newly created and, thus, has a higher surface 331

FIG. 4. The maximum height of the Rayleigh jet for SDS, CTAB, and Brij-35. FIG. 5. Initial speed of the Rayleigh jet for SDS, CTAB, and Brij-35.

Phys. Fluids 34, 000000 (2022); doi: 10.1063/5.0098642 34, 000000-5


Published under an exclusive license by AIP Publishing

ID: aipepub3b2server Time:


ID: aipepub3b2server Time: 18:05
12:22 II Path:
Path: D:/AIP/Support/XML_Signal_Tmp/AI-APL#210696
D:/AIP/Support/XML_Signal_Tmp/AI-PHF#221109
J_ID:
J_ID:APPLAB
PHFLE6 DOI:
DOI:10.1063/5.0045281
10.1063/5.0098642 Date:
Date:8-April-21
2-July-22 Stage:
Stage: Page:
Page: 76 Total
Total Pages:
Pages: 89

PROOF
PROOF COPY
COPY [APL21-AR-00653]
[POF22-AR-02572]

Applied Physics
Physics of FluidsLetters ARTICLE
ARTICLE scitation.org/journal/apl
scitation.org/journal/phf

liquid. The dynamic surface tension of the jet is estimated with 338
dynamic surface tension measurements at a time of 90 ms, which is 339
the characteristic time of the jet formation. Both surface tensions are 340
plotted in Fig. 8 at different concentrations for the three tested surfac- 341
tants. The equilibrium surface tension displays the expected behavior, 342
i.e., a decrease in the concentration until stabilization at the CMC. 343
However, the dynamic surface tension displays a different behavior, 344
with almost no changes for low concentrations, then a decrease with 345
the concentration. This offset between the two curves leads to a rise 346
and a decrease in the surface-tension difference with a maximum at 347
the same concentrations as the minimum of acceleration. Therefore, 348
the higher the surface tension difference, the lower the acceleration. 349
Indeed, the forces induced by the Marangoni flow act against the grav- 350
ity and surface tension forces, lowering the acceleration undergone by 351
the jet. Indeed, Fig. 9 shows the value of acceleration as a function of 352
the surface-tension difference for the three surfactants. A clear linear 353
law is visible, showing that the Marangoni flows are mainly responsi- 354
ble for the acceleration behavior. 355
Moreover, how to explain this difference in behavior between the 356
three surfactants, and what would be the physicochemical origin of 357
this dynamic? The keying would be the surfactant charges. Indeed the 358
FIG. 6. Rayleigh jet acceleration during its rise for SDS, CTAB, and Brij-35. osmotic pressure of counter ions is significantly higher than for neutral 359
surfactants and could contribute to a faster homogenization of the sur- 360
332 tension than the rest of the surface, so the Marangoni flows are factant concentration for charged ones. Therefore, the concentration 361
333 directed in the same way as the jet flows. A similar phenomenon gradients of SDS and CTAB being less important than for Brij-35 362
334 involving Marangoni flows is reported for water drop impact on etha- would lead to higher jet initiation velocities for the latter. Indeed, an 363
335 nol.31 The Marangoni flow speed is proportional to the surface tension indirect measure of the equilibration effect of osmotic pressure can be 364
336 difference between the higher and lower points. The surface tension of read from Fig. 8. The latter shows the difference between the 365
337 the liquid at rest corresponds to the equilibrium surface tension of the

FIG. 7. Maximum height of the Rayleigh jet as a function of jet initial speed and
acceleration for SDS, CTAB, and Brij-35. The filled and unfilled symbols correspond
to concentrations below and above the CMC, respectively. The grayed-out points
represent cases when a secondary droplet detaches during the jet rise. The black FIG. 8. Equilibrium (open symbols) and dynamic (close symbols) surface tensions
line represents Eq. (2). of SDS, CTAB, and Brij-35 as a function of the fraction of the CMC.

Phys. Fluids 34, 000000 (2022); doi: 10.1063/5.0098642 34, 000000-6


Published under an exclusive license by AIP Publishing

ID: aipepub3b2server Time:


ID: aipepub3b2server Time: 18:05
12:22 II Path:
Path: D:/AIP/Support/XML_Signal_Tmp/AI-APL#210696
D:/AIP/Support/XML_Signal_Tmp/AI-PHF#221109
J_ID:
J_ID:APPLAB
PHFLE6 DOI:
DOI:10.1063/5.0045281
10.1063/5.0098642 Date:
Date:8-April-21
2-July-22 Stage:
Stage: Page:
Page: 77 Total
Total Pages:
Pages: 89

PROOF
PROOF COPY
COPY [APL21-AR-00653]
[POF22-AR-02572]

Applied Physics
Physics of FluidsLetters ARTICLE
ARTICLE scitation.org/journal/apl
scitation.org/journal/phf

numerical simulations of drop impact could be made by taking into 396


account the surfactant dynamics. 397

SUPPLEMENTARY MATERIAL 398

See the supplementary material to distinguish different groups of 399


acoustic events presented in Fig. 1 in addition to the variation of the 400
maximum height of the Rayleigh jet as a function of surface tension 401
for aqueous solutions of ethanol. 402
403
ACKNOWLEDGMENTS 404

The authors would like to express their sincere thanks to 405


Mr. James Blondeau, assistant laboratory engineer at LAUM, who 406
efficiently and significantly contributed to the setup of the 407
experiments. Our thanks also go to La Region des Pays de la Loire, 408
which financially supported this work within the framework of the 409
Acapulcos project of the RFI LMAC. 410
411
AUTHOR DECLARATIONS 412
Conflict of Interest 413

FIG. 9. Rayleigh jet acceleration as a function of the difference between dynamic The authors have no conflicts to disclose. 414
and equilibrium surface tensions for SDS, CTAB, and Brij-35.
Author Contributions 415

Gautier Gillot: Data curation (lead); Formal analysis (equal); 416


366 equilibrium and “dynamic” surface tension estimated over a measure-
Investigation (lead); Methodology (equal); Software (lead); Writing – 417
367 ment time equivalent to the duration of the hydrodynamic/acoustic
original draft (equal); Writing – review and editing (equal). Jean- 418
368 event probed in this work (90 ms). This measurement clearly shows
Michel Genevaux: Funding acquisition (equal); Methodology 419
369 a larger deviation between the two-surface tension values for Brij-35
(supporting); Project administration (supporting); Supervision (sup- 420
370 than for SDS and CTAB.
porting); Validation (equal); Writing – review and editing (equal). 421
371 CONCLUSION Laurent Simon: Funding acquisition (equal); Methodology (equal); 422
Project administration (supporting); Software (supporting); Validation 423
372 The effect of a surfactant on the hydrodynamics and acoustics of
(equal); Writing – review and editing (equal). Lazhar Benyahia: 424
373 a water drop impacting a pool of aqueous solutions is studied. A uni-
Conceptualization (lead); Data curation (equal); Formal analysis 425
374 versal behavior is observed concerning the evolution of the maximum
(lead); Funding acquisition (lead); Methodology (lead); Project admin- 426
375 height of the Rayleigh jet for three different surfactants, when the sur-
istration (lead); Resources (equal); Supervision (lead); Validation 427
376 factant concentration is normalized by its CMC. In particular, the
(lead); Writing – original draft (lead); Writing – review and editing 428
377 maximum height of the Rayleigh jet shows a maximum around the
(lead). 429
378 CMC/3. This feature is also observed at the starting time of the acous-
430
379 tic events where a maximum is also observed around CMC/3. This
380 universal behavior is related to the surfactant dynamics as expressed in DATA AVAILABILITY 431
381 the Marangoni flows at the creation of a new surface following the for- The data that support the findings of this study are available 432
382 mation of the jet. from the corresponding authors upon reasonable request. 433
383 On the one hand, the variation of the jet height is linked to the
384 increase in viscosity above the CMC. On the other hand, the presence
REFERENCES 434
385 of Marangoni flows takes place because of the surface-tension gradient 1
G. J. Franz, “Splashes as sources of sound in liquids,” J. Acoust. Soc. Am. 31, 435
386 between the bath surface at rest and the newly created surface of the 436
1080 (1959).
387 jet. This phenomenon is responsible for the acceleration of the jet and 2
Y. D. Chashechkin and V. E. Prokhorov, “Acoustics and hydrodynamics of a 437
388 allows us to understand the subtle differences between the surfactants. drop impact on a water surface,” Acoust. Phys. 63, 33 (2017). 438
3
389 Note that the presence of Marangoni flows has already been observed H. C. Pumphrey and P. A. Elmore, “The entrainment of bubbles by drop 439
390 in the case of drop impact 31 but was not related to the acoustic of the 4
impacts,” J. Fluid Mech. 220, 539 (1990). 440
R. Chicharro, R. Manasseh, and A. Vazquez, “The heart signal: An acoustic sig- 441
391 impact nor to the physicochemical properties or the adsorption kinet-
nature observed during a second-bubble entrainment,” Chem. Eng. Sci. 219, 442
392 ics of surfactants. The acoustic signature of the drop impact is 443
115597 (2020).
393 revealed to be a new and efficient tool to investigate the CMC of the 5
Y. Tomita, T. Saito, and S. Ganbara, “Surface breakup and air bubble forma- 444
394 surfactant. It could also help to better understand the surfactant fast tion by drop impact in the irregular entrainment region,” J. Fluid Mech. 588, 445
395 dynamics.32–34 In order to quantify more precisely these results, 131 (2007). 446

Phys. Fluids 34, 000000 (2022); doi: 10.1063/5.0098642 34, 000000-7


Published under an exclusive license by AIP Publishing

ID: aipepub3b2server Time:


ID: aipepub3b2server Time: 18:05
12:22 II Path:
Path: D:/AIP/Support/XML_Signal_Tmp/AI-APL#210696
D:/AIP/Support/XML_Signal_Tmp/AI-PHF#221109
J_ID:
J_ID:APPLAB
PHFLE6 DOI:
DOI:10.1063/5.0045281
10.1063/5.0098642 Date:
Date:8-April-21
2-July-22 Stage:
Stage: Page:
Page: 78 Total
Total Pages:
Pages: 89

PROOF
PROOF COPY
COPY [APL21-AR-00653]
[POF22-AR-02572]

Applied Physics
Physics of FluidsLetters ARTICLE
ARTICLE scitation.org/journal/apl
scitation.org/journal/phf

447 6 20 488
G. Gillot, C. Derec, J. M. Genevaux, L. Simon, and L. Benyahia, “A new insight N. M. Kovalchuk and M. J. H. Simmons, “Surfactant-mediated wetting and
448 on a mechanism of airborne and underwater sound of a drop impacting a liq- spreading: Recent advances and applications,” Curr. Opin. Colloid Interface 489
449 uid surface,” Phys. Fluids 32, 062004 (2020). Sci. 51, 101375 (2021). 490
450 7 21 491
E. Castillo-Orozco, A. Davanlou, P. K. Choudhury, and R. Kumar, “Droplet K. Kinoshita, E. Parra, and D. Needham, “Adsorption of ionic surfactants at
451 impact on deep liquid pools: Rayleigh jet to formation of secondary droplets,” microscopic air-water interfaces using the micropipette interfacial area- 492
452 Phys. Rev. E 92, 053022 (2015). expansion method: Measurement of the diffusion coefficient and renormaliza- 493
453 8 494
F. Marcotte, G. J. Michon, T. Seon, and C. Josserand, “Ejecta, corolla, and tion of the mean ionic activity for SDS,” J Colloid Interface Sci. 504, 765
454 splashes from drop impacts on viscous fluids,” Phys. Rev. Lett. 122, 014501 (2017). 495
455 22 496
(2019). K. Kinoshita, E. Parra, and D. Needham, “New sensitive micro-measurements
456 9 497
G.-J. Michon, C. Josserand, and T. Seon, “Jet dynamics post drop impact on a of dynamic surface tension and diffusion coefficients: Validated and tested for
457 deep pool,” Phys. Rev. Fluids 2, 023601 (2017). the adsorption of 1-Octanol at a microscopic air-water interface and its disso- 498
458 10 499
A. Ogawa, K. Utsuno, M. Mutou, S. Kouzen, Y. Shimotake, and Y. Satou, lution into water,” J Colloid Interface Sci. 488, 166 (2017).
459 23 500
“Morphological study of cavity and Worthington jet formations for Newtonian J. N. Israelachvili, Intermolecular and Surface Forces (Academic Press, London,
AQ2 460 and non-Newtonian liquids,” Part. Sci. Technol. 24, 181 (2006). 2011). 501
461 11 24 502
C. Tang, “An experimental study on droplet impact on a deep liquid pool: R. Zana, “Dimeric and oligomeric surfactants. Behavior at interfaces and in
462 impact outcomes and their transition boundaries,” J. Water Sci. Eng. 1, 䊏 aqueous solution: review,” Adv. Colloid Interface Sci. 97, 205 (2002). 503
25
AQ3 463 (2020). Y. Ephraim and D. Malah, “Speech enhancement using a minimum-mean 504
464 12 505
J. M. Cheny and K. Walters, “Rheological influences on the splashing square error short-time spectral amplitude estimator,” IEEE Trans. Acoust.,
465 experiment,” J. Non-Newtonian Fluid Mech. 86, 185 (1999). Speech, Signal Process. 32, 1109 (1984). 506
13 26
466 A. M. Karim, “Experimental dynamics of Newtonian non-elastic and visco- V. Hella, “Digital audio restoration,” M.S. thesis (䊏, 2013). 507 AQ4
467 27 508
elastic droplets impacting immiscible liquid surface,” AIP Adv. 9, 125141 C. J. M. van Rijn, “Emanating jets as shaped by surface tension forces,”
468 (2019). Langmuir 34, 13837 (2018). 509
14 28
469 A. M. Karim, “Experimental dynamics of Newtonian and non-Newtonian H. Ma, C. Liu, X. Li, H. Huang, and J. Dong, “Deformation characteristics and 510 AQ5
470 droplets impacting liquid surface with different rheology,” Phys. Fluids 32, energy conversion during droplet impact on a water surface,” Phys. Fluids 31, 511
471 043102 (2020). 062108 (2019). 512
472 15 29 513
T. Q. Nguyen-Pham, L. Benyahia, G. Bastiat, J. Riou, and M. C. Venier- H. Akbas, T. Sidim, and M. Iscan, “Effect of polyoxyethylene chain length and
473 Julienne, “Behavior of poly(D,L-lactic-co-glycolic acid) (PLGA)-based droplets electrolyte on the viscosity of mixed micelles,” Turk. J. Chem. 27, 357 (2003). 514
474 30 515
falling into a complex extraction medium simulating the prilling process,” J. M. Khademi, W. Wang, W. Reitinger, and D. P. J. Barz, “Zeta Potential of
475 Colloid Interface Sci. 561, 838 (2020). potential of poly(methyl methacrylate) (PMMA) in contact with aqueous 516
476 16 517
H. C. Pumphrey, L. A. Crum, and L. Bjo/rno, “Underwater sound produced electrolyte-surfactant solutions,” Langmuir 33, 10473 (2017).
477 31 518
by individual drop impacts and rainfall,” J. Acoust. Soc. Am. 85, 1518 F. Jia, K. Sun, P. Zhang, C. Yin, and T. Wang, “Marangoni effect on the impact
478 (1989). of droplets onto a liquid-gas interface,” Phys. Rev. Fluids 5, 073605 (2020). 519
479 17 32 520
G. Gillot, L. Simon, J. M. Genevaux, and L. Benyahia, “Acoustic signatures and Y. Han, J. Koplik, and C. Maldarelli, “Surfactant and dilatational viscosity
480 bubble entrainment mechanisms of a drop impacting a water surface with effects on the deformation of liquid droplets in an electric field,” J Colloid 521
481 surfactant,” Phys. Fluids 33, 077114 (2021). Interface Sci. 607, 900 (2022). 522
482 18 33 523
M. Kalli, L. Chagot, and P. Angeli, “Comparison of surfactant mass transfer B. Petkova, S. Tcholakova, M. Chenkova, K. Golemanov, N. Denkov, D.
483 with drop formation times from dynamic interfacial tension measurements in Thorley, and S. Stoyanov, “Foamability of aqueous solutions: Role of surfactant 524
484 microchannels,” J. Colloid Interface Sci. 605, 204 (2022). type and concentration,” Adv. Colloid Interface Sci. 276, 102084 (2020). 525
485 19 34 526
N. M. Kovalchuk, H. Jenkinson, R. Miller, and M. J. H. Simmons, “Effect of sol- J. Zhou and P. G. Ranjith, “Insights into interfacial behaviours of surfactant
486 uble surfactants on pinch-off of moderately viscous drops and satellite size,” J and polymer: A molecular dynamics simulation,” J. Mol. Liq. 346, 117865 527
487 Colloid Interface Sci. 516, 182 (2018). (2022). 528

Phys. Fluids 34, 000000 (2022); doi: 10.1063/5.0098642 34, 000000-8


Published under an exclusive license by AIP Publishing

ID: aipepub3b2server Time:


ID: aipepub3b2server Time: 18:05
12:22 II Path:
Path: D:/AIP/Support/XML_Signal_Tmp/AI-APL#210696
D:/AIP/Support/XML_Signal_Tmp/AI-PHF#221109

You might also like