23
Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefiane a, , Charles A. Ward b,1 a School of Engineering and Electronics, The University of Edinburgh, The Kings Buildings, Mayfield Road Edinburgh, EH9 3JL United Kingdom b Department of Mechanical and Industrial Engineering, University of Toronto, 5 King's College Road, Toronto, Canada M5S 3G8 Available online 5 May 2007 Abstract Thermocapillary convection has a very different history for water than for other liquids. For water, several studies have pointed to the lack of evidence supporting the existence of thermocapillary (or Marangoni) convection. Other studies have given clear evidence of its existence and of the role it plays during steady-state water evaporation. We examine both sets of data and suggest a reason for the difference in the interpretation of the experimental data. For organic liquids, the evidence of thermocapillary convection has been clearly documented, but the issues are the type of flow that it generates during steady-state evaporation. We review the measurements and show that the flow field of the evaporating liquid is strongly affected by the presence of the thermocapillary convection. When the results obtained from both water and organic liquids are compared, they give further insight into the nature of thermocapillary convection. © 2007 Elsevier B.V. All rights reserved. Contents 1. Introduction ............................................................. 201 2. Surface tension-driven convection in water ............................................. 202 2.1. Does surface tension-driven convection exist in water? ................................... 203 2.2. The energy criterion for the onset of surface tension-driven convection ........................... 205 2.3. The transition to interfacial turbulence during water evaporation ................................. 207 2.4. Interfacial temperature discontinuity during steady-state evaporation ............................. 207 3. Interfacial driven flows during evaporation of non-polar liquids ................................... 209 4. Interfacial conditions and thermocapillary flows in organic liquids ................................. 213 4.1. Spontaneously evaporating non-polar liquids in capillary tubes ............................... 213 4.2. Receding evaporating meniscus and observed flows ..................................... 214 4.3. Temperature distribution ................................................... 217 4.4. Fluid flow measurements ................................................... 217 4.5. Imposed external heating and reversal of the flow ...................................... 218 5. Discussion .............................................................. 220 References ................................................................. 223 1. Introduction Molecules in a liquid experience attractive, intermolecular van der Waals forces. These forces are caused by the arrangement of charge distribution inside molecules. Therefore, in a body of liquid the net force on a particle will be zero due to Advances in Colloid and Interface Science 134135 (2007) 201 223 www.elsevier.com/locate/cis Corresponding author. Tel.: +44 131 650 4873; fax: +44 131 650 6551. E-mail addresses: [email protected] (K. Sefiane), [email protected] (C.A. Ward). 1 Tel.: +1 416 978 4807; fax: +1 416 978 7322. 0001-8686/$ - see front matter © 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.cis.2007.04.020

Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Science 134–135 (2007) 201–223www.elsevier.com/locate/cis

Advances in Colloid and Interface

Recent advances on thermocapillary flows and interfacial conditionsduring the evaporation of liquids

Khellil Sefiane a,⁎, Charles A. Ward b,1

a School of Engineering and Electronics, The University of Edinburgh, The Kings Buildings, Mayfield Road Edinburgh, EH9 3JL United Kingdomb Department of Mechanical and Industrial Engineering, University of Toronto, 5 King's College Road, Toronto, Canada M5S 3G8

Available online 5 May 2007

Abstract

Thermocapillary convection has a very different history for water than for other liquids. For water, several studies have pointed to the lack ofevidence supporting the existence of thermocapillary (or Marangoni) convection. Other studies have given clear evidence of its existence and ofthe role it plays during steady-state water evaporation. We examine both sets of data and suggest a reason for the difference in the interpretationof the experimental data. For organic liquids, the evidence of thermocapillary convection has been clearly documented, but the issues are the typeof flow that it generates during steady-state evaporation. We review the measurements and show that the flow field of the evaporating liquid isstrongly affected by the presence of the thermocapillary convection. When the results obtained from both water and organic liquids are compared,they give further insight into the nature of thermocapillary convection.© 2007 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2012. Surface tension-driven convection in water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

2.1. Does surface tension-driven convection exist in water? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2032.2. The energy criterion for the onset of surface tension-driven convection . . . . . . . . . . . . . . . . . . . . . . . . . . . 2052.3. The transition to interfacial turbulence during water evaporation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2072.4. Interfacial temperature discontinuity during steady-state evaporation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207

3. Interfacial driven flows during evaporation of non-polar liquids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2094. Interfacial conditions and thermocapillary flows in organic liquids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213

4.1. Spontaneously evaporating non-polar liquids in capillary tubes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2134.2. Receding evaporating meniscus and observed flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2144.3. Temperature distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2174.4. Fluid flow measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2174.5. Imposed external heating and reversal of the flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218

5. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223

⁎ Corresponding author. Tel.: +44 131 650 4873; fax: +44 131 650 6551.E-mail addresses: [email protected] (K. Sefiane), [email protected]

(C.A. Ward).1 Tel.: +1 416 978 4807; fax: +1 416 978 7322.

0001-8686/$ - see front matter © 2007 Elsevier B.V. All rights reserved.doi:10.1016/j.cis.2007.04.020

1. Introduction

Molecules in a liquid experience attractive, intermolecularvan der Waals forces. These forces are caused by thearrangement of charge distribution inside molecules. Therefore,in a body of liquid the net force on a particle will be zero due to

Page 2: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

202 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

the deleterious effect of the forces exerted by the neighbouringmolecules. However, at the surface of a liquid there is a net forceacting inwards because there is no attractive force from thevapour side of the molecules. As a result, the molecules at thesurface will contract in towards the body of liquid, and willresist penetrations; this effect is called surface tension.

In 1855 Tomson [1] investigated surface tension effects indrops of alcohol spreading on a water surface; however his workwent unmentioned till more than a decade later, Marangoni [2]contested the explanations given by Van der Mensbrugghe [3] tothese observations. Carlo Giuseppe Matteo Marangoni (1840–1925) [4], studied the effects of one liquid spreading uponanother. His findings, published in 1871, proved that a givenliquid A, spreads on another liquid B, when the sum of theinterfacial tension and the surface tension of A, is lower than thesurface tension of B. At the start of the 20th Century, Bénard[5,6] observed cellular convective motion in thin layers of aliquid above heated surfaces. These observations were put downto Rayleigh convection, or buoyancy, until J. Pearson [7] showedthat the motion could also be induced by the effects of surfacetension, as had originally been proposed by Marangoni [2]. Hisinterest was aroused by experiments in which it was noted thatdrying paint films often displayed steady cellular circulatoryflow similar to that seen in thin liquid layers heated from below.The latter can be explained by the effects of buoyancy; howeverthe former cannot because the convection cells occur whetherthe free surface is on the top or the bottom of the paint layer. Thecause of the flows was explained as being due to surface tensionwhich, for most fluids at most temperatures, is a decreasingfunction of temperature (and if the mixture is a binary one, afunction of concentration). Therefore, if the surface is not atuniform temperature (or concentration), surface tension varia-tions arise which drive the flows of the fluids. The explanationproposed by Pearson attributes the cause of the flow to surfacetension effect. Pearson also showed that for the convection cellsto form under the action of surface tension forces certainrequirements have to be met.

Marangoni convection is a flow phenomenon at liquid–vapour or liquid–liquid interfaces due to a gradient in surfacetension. These gradients can be brought about by variations inthe liquid concentration or temperature and result in solutal-Marangoni convection or thermocapillary convection respec-tively. The non-uniformities lead to variations in the surfacetension force along the interface and flow of the liquid along theforce gradient. Owing to the fact that the surface instabilitiesaffect local regions only, the liquid flows that are generated canform a cellular pattern which leads to enhanced mass and heattransfer in the vicinity of the interface. These capillary-drivenfluid flows have attracted increasing attention of the scientificcommunity during the last century. These induced flows can beof a paramount importance in a wide range of biological andindustrial areas. Although literature reviews of this area haveappeared (Levich [8,9], Kenning [10], Davis [11], Normand[12], Velarde [13] Schatz et al. [14]), recent advances intechnological applications of these phenomena have stressedthe need for better understanding. To that end, we divide ourconsiderations into those for water and those for organic liquids.

At around about the same time as Pearson [7], Sternling andScriven [15,16] published a paper considering the influence ofthe Marangoni effect on turbulence at liquid–liquid interfaces.They considered the case of two liquid phases in contact witheach other and undergoing mass transfer by diffusion, andproposed that if a disturbance arises at the interface it coulddisrupt the convection of fluid, with the result dependent on therelative kinematic viscosities and diffusivities of the liquids.

If this leads to a variation in concentration at the interface it willproduce corresponding variations in the interfacial tension. As aresult the interface is no longer in equilibrium and so will try tomove to a state of lower free energy through expansion of regionsof low interfacial tension at the expense of those regions withhigher interfacial tension. Thismeans that because there cannot bediscontinuities in the velocities of the liquids at the interface,motion in one region of the liquid induces motion in itsneighbours. Therefore, Sternling and Scriven [15] introduce theeffects of the solutal-Marangoni effect on liquid layers. This effectis also discussed by Velarde [17] who mentions that surfacetension-driven flows can arise under isothermal conditions if, in abinary mixture, the component of lower surface tension or lowerdensity is desorbed into another phase. Alternatively, the sameresult could be produced by heating non-evaporating liquids frombelow and allowing them to cool at the top. Lucassen [18,19] hasadopted a different approach in the analysis of Marangoni flows;the interface is considered as a membrane which can propagatelongitudinal and transversal waves. Based on the above approachand using linear stability analysis, Gouda [20] predicted the onsetof Marangoni convection in a consistent way with the approachdeveloped by Sternling and Scriven [15]. Many other studiescontinued to use the concept of longitudinal waves, called bysome, Lucassen–Marangoni waves.

In what follows the results from two independent investiga-tions on the evaporation of liquids and thermocapillary flow willbe presented. Both studies are experimental works looking in therole of interfacial conditions in driving thermocapillary flowsduring evaporation of liquids. The first study investigates steadyevaporation of water under reduced pressure. Temperatureprofiles in liquid and vapour near the interface aremeasured. Thefluid flow in the liquid phase is investigated using a probe toreveal the orientation and strength of convection. The secondstudy investigates the evaporation of organic liquids of variousvolatilities in capillary tubes of different sizes. The investigationexamines the existence and nature of fluid flow in the liquidphase during evaporation of these liquids. Despite thedifferences between the two studies the conclusions reachedare corroborating each other. Some interesting analogiesbetween the conclusions reached by both studies can be drawn.

2. Surface tension-driven convection in water

Pearson [7] considered a horizontal liquid layer heated frombelow in which the temperature gradient was constant and askedwhen the system would become unstable. He neglectedevaporation, and with this approximation found the liquidremained motionless until a non-dimensional number that laterbecame called the Marangoni number, Ma exceeded ~80. If the

Page 3: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 1. Schematic of a stainless-stainless funnel used to study the temperaturefield in the liquid and vapour phases of water as it evaporated under steady-stateconditions [26].

203K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

liquid-layer thickness is denoted as D, the surface tension asγLV, the thermal diffusivity as αL, the dynamic viscosity by η,Ma may be expressed

Ma ¼ AgLV

AT

� �1

ðT0 � TIÞDaLg

ð1Þ

where TI is the interfacial temperature, and T0 is the temperatureat the bottom of the liquid layer. If the experimental circumstanceswere such that Ma exceeded this threshold value, Pearson'sprediction was that convection along the interface would ensue.For organic liquids, this criterion for the on-set of surface tension-driven convection has received significant support (see below),but for water the results have been less clear.

2.1. Does surface tension-driven convection exist in water?

The existence of surface tension-driven convection for waterhas been controversial; we first review the evidence both againstand for its existence. Barnes, Hunter and Feher, and Schreiber,Cammenga and Rudolph [22–25] have reported experimentswith carefully cleaned water in which Ma was greater than thethreshold value, but they did not detect surface tension-drivenconvection. Their basic criterion was to measure the temperatureprofile in the liquid phase near the interface, and if the temperatureprofile were linear, they concluded the energy transport was bythermal conduction, and there was no surface tension-drivenconvection at the interface.Aswill be seen, the critical issue is howclose to the interface they were able to measure the temperature.

These investigators used different techniques to establish thetemperature profile during evaporation. Barnes and Feher [22]and Barnes and Hunter [21] used interferometry. Their first datapoint was ∼0.125 mm below the interface, and their second onewas 0.625 mm below. The temperature profile between thesepoints was assumed linear. But between 0.125 mm and theinterface, they were unable to measure the temperature profile.Schreiber, Cammenga and Rudolph [23,24] used thermometry tomeasure the temperature profile, but they do not state how close tothe interface they were able to measure the water temperature.However, the water–vapour interface was moving, due to theevaporation, during their measurements, and this must haveintroduced some uncertainty in the position at which temperaturewas measured. Cammenga et al. [25] summarized the results ofthe experiments by the two experimental groups. After noting thatBarnes and Hunter had not observed surface tension-drivenconvection in their experiments, it was pointed out that this “isconsistent with the results of Schreiber and Cammenga”.

Another approach to studying the temperature near theinterface of water as it evaporates was introduced by Ward andDuan [26]. The stainless-steel funnel shown schematically inFig. 1 was enclosed in a chamber where the pressure could becontrolled. During each experiment, water was supplied to thefunnel throat with a syringe pump at the same rate as the waterevaporated at the funnel mouth; thereby, the interface was kept ata constant height. The temperature at the funnel throat could becontrolled by a heat bath and the water temperature there could bemeasured with a thermocouple permanently placed at the throat.

A second thermocouple constructed of 25 μm diameter type Kthermocouple wire [27] was formed into a U-shape. The ratio ofthe bottom-length of the U-shape to the wire diameter was morethan 20 [28]. The thermocouple was mounted on a positioningmicrometer that allowed the thermocouple to be positioned withan accuracy of ±0.01 mm. As the water evaporated under steady-state conditions, the water temperature was measured as afunction depth, starting at 0.05±0.01 mm below the interface andmoving down in steps of 0.02 mm to a maximum depth of 1 mm[29].

In one series of experiments, the temperature at the funnelthroat wasmaintained at∼3.5 °C, and the water–vapour interfaceon the funnel centreline 1 mm above the funnel mouth. When theheight of the liquid–vapour interface was not more than 1.0 mm,the interface could be approximated as spherical and its positioncould be measured with a cathetometer. The experimental pro-cedure did not allow the interface to move by more than ±10 μmduring the course of the experiment. The syringe pump suppliedwater at a steady rate, and the pumping rate could be accuratelymeasured by observing the piston with a cathetometer. Thepressure in the chamber and the pumping rate were adjusted so thewater–vapour interface did not move. Under this condition, thepumping rate is equal to the evaporation rate.

The temperaturemeasured on the funnel centreline as thewaterevaporated at a rate of 3.38±0.03 g/m2 s is shown in Fig. 2 [26].As seen there: (1) the measured temperature was strictly uniformto a depth of approximately 0.075 mm, and below 0.125 mm, thetemperature gradient was very nearly constant. (2) A temperaturediscontinuity of 4 °C was measured at the interface in which theinterfacial vapour temperature was greater than that of the liquid.This temperature discontinuity is discussed in Section 2.3.

Except for the uniform temperature layer and the temperaturediscontinuity, these results are consistent with results reportedby Barnes et al. [21,22], but since they were not able to measurethe temperature closer to the interface than 0.125 mm with theinterferometer technique, they did not observe the uniform-temperature layer. The uniform-temperature layer is suggestiveof surface tension-driven convection being present. They didnot report temperature measurements in the vapour phase.

Page 4: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 3. The upper graph shows the measured temperature profile parallel to theinterface of water as it evaporated at 3.38±0.03 g /m2s into its vapour main-tained at 300 Pa [26]. The temperature measured as a function of depth on thecentreline is shown in Fig. 2. The lower graph shows the depth of the uniformtemperature layer.

Fig. 2. The measured temperature [26] on the centreline of water maintained atthe mouth of the funnel shown in Fig. 1 while evaporating at 3.38±0.03 g /m2s.The vapour phase pressure was maintained at 300 Pa.

204 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

The driving-potential for this convection was observed whenthe temperature parallel to the interface was measured. Thisresult is seen in the upper graph in Fig. 3. The temperature fieldwas approximately axi-symmetric, and the difference in theinterfacial liquid temperature between the funnel rim and thecentreline was less than approximately 0.2 °C. In the lowergraph of this figure, the uniform-temperature layer is seen tohave varied along the interface: its maximum value was on thecentreline and minimum at the funnel rim.

Since the temperature at the funnel throat corresponded withthat at which water has its maximum density, and the evaporationat the liquid–vapour interface lowers the interfacial temperaturebelow the throat temperature, on the centreline the lighter liquidwas above the heavier. Parallel to the interface, the same conditionprevails: the interfacial liquid temperature at the funnel rim washigher than that on the centreline (see Fig. 3), but since bothtemperatures are below 4 °C, the water density at the funnel rim isgreater than that on the centreline, and since the interface wasspherical in shape, the lighter liquid is above the heavier. Thus,throughout the funnel there could have been no buoyancy-drivenconvection, and the explanation for the uniform-temperature layermust be sought elsewhere. We note that a uniform-temperaturebelow the water–vapour interface during both water evaporationand condensation has been reported before [30].

Eq. (1) may be applied to determine if the criterion proposedby Pearson indicates a linear temperature profile would beunstable. As indicated in Fig. 2, when the liquid was evapo-rating under steady state conditions, the temperature profilebelow the uniform-temperature layer was very nearly linear. Ifthis temperature profile is extrapolated to the interface to defineTI, and D in Eq. (1) is taken to be the distance from the throat tothe interface on the centreline, one finds the value of Ma to be

more than 21,000 [29]. Thus, the Pearson criterion wouldcertainly suggest the linear temperature profile would beunstable and surface tension-driven convection would result.This convection would be a possible explanation for theobserved uniform-temperature layer, but to determine if theconvection was in-fact present, a series of experiments wasconducted with a flow probe.

The same apparatus as that of Ward and Duan [26] was used,except the thermocouple was replaced with a “flow probe”: a12.5 μm-diameter cylinder, 52 mm in length. One end of theflow probe was cantilever mounted on the positioningmicrometer. The other end could be inserted in the liquid. Theseries of experiments that had been run to determine thetemperature profile during steady-state water evaporation wasrepeated. In each experiment, the throat temperature andpressure in the vapour phase were maintained at values thatcorresponded to an experiment run to determine the temperatureprofile. After the water was evaporating steadily, the probe tipwas lowered until it touched the liquid surface. The probe wasobserved to immediately deflect toward the funnel centreline[26]. It was then inserted approximately 40 μm into the liquid,and its position on the interface was recorded with a videocamera mounted on a telescope. A photograph of the probewhen it was in the liquid phase is shown in Fig. 4. Also shownin this figure is the amplitude of the probe oscillations as afunction of time. They were determined from the results recor-ded by the video camera.

In experiment EV10, there was no measurable evaporation:the measured pressure in the vacuum chamber was greater thanthe saturation vapour pressure corresponding to the measuredtemperature in the funnel throat [26]. In experiment EV11, theaverage evaporation flux was 0.407 g/m2 s;Mawas∼2200; andas indicated in Fig. 4, the probe detected an oscillatory flowthat was on average directed toward the funnel centreline. Whenthe evaporation rate was raised further (EV15: 2.544 g/m2 s;Ma 18,200), the probe deflection increased, but so did the mag-nitude of the probe oscillations.

Page 5: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 5. Values of the fluid speed parallel to the interface measured with the flowprobe, and independently from the measured temperature gradient parallel to theinterface [26].

Fig. 4. The upper photograph shows the 12.5 µm-diameter probe that wasinserted to a depth of 40 µm into the liquid phase as the liquid evaporated whilemaintained at the mouth of a stainless-steel funnel. The probe was observed todeflect and oscillate about its deflected position. The amplitudes of the probeoscillations are shown in the lower graph [26].

205K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

The deflection of the flow probe supports the hypothesis thatthere was surface tension-driven convection in the waterevaporation experiments when Ma was greater than 2200, butit does not indicate that the surface tension-gradient resultingfrom the measured temperature gradient parallel to the interfacecaused the flow that produced the probe deflection. Todetermine if the measured temperature gradient parallel to theinterface, such as that shown in Fig. 3, could be responsible forthe probe deflection, calculations were made in two ways.

Under steady state conditions, no acceleration of the flow atthe interface would be expected; thus the shear stress parallel tothe interface σrθ would be balanced by the surface tensiongradient. In spherical coordinates (R, θ, ϕ)

rrhÞr¼R ¼ ðjgLVÞr¼Rd ih ð2Þ

The measured interfacial liquid temperature was approxi-mately axi-symmetric, and Ward and Duan [26] neglected thedependence of the radial speed, vr on the polar angle, θ. Thisallowed them to integrate Eq. (2) and construct an expressionfor the speed parallel to the interface, vθ in terms of thethickness of the uniform-temperature layer thickness, δu and(dTI

L / dθ)

vh ¼ � 1gL

� �dgLV

dTLI

� �dTL

I

dh

� �ln 1� du

R0

� �� �ð3Þ

The integration constant was evaluated by assuming vθvanished at the maximum depth of the uniform-temperaturelayer. When the measured thickness of the uniform-tempera-ture layer (see Fig. 3) and the measured gradient in theinterfacial liquid temperature were used in Eq. (3), the valuesof vθ indicated in Fig. 5 were obtained. This is the fluid speedthat is predicted to result from the measured temperaturegradient parallel to the interface. The speed calculated fromEq. (3) is the mean fluid speed. The flow probe indicated anoscillatory flow.

In the second calculation method, the fluid speed that wouldgive a deflection ymax to the probe tip was estimated: if theexpression for the drag on a cylinder obtained from the Oseenapproximation for flow of an incompressible liquid directedperpendicularly to the longitudinal axis of the cylinder [29] isequated to the expression for the force required to deflect the tipof a cantilever-mounted “beam” a distance ymax, one finds anexpression that can be solved iteratively for the flow speed [26].The values of the speed determined in this way are also shownin Fig. 5. Note that the mean of the fluctuating speeds –calculated from the probe deflections – corresponds reasonablywith those calculated from the measured temperature gradient,Eq. (3). This suggests that the convection observed with theprobe results from the surface-tension gradient produced by thetemperature gradient.

2.2. The energy criterion for the onset of surface tension-drivenconvection

In the experiments reported in References [26] and [29], theevaporation rate was increased by lowering the pressure in the

Page 6: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 7. Comparison of the energy transport required to evaporate the liquid at themeasured rate with the energy transport to the interface by thermal conduction.

Fig. 6. The value of the Marangoni number as a function of the averageevaporation flux [26,29].

206 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

vapour while maintaining the temperature at the funnel throat atapproximately the same value, ∼3.5 °C.

At the highest pressures (lowest evaporation rate), thesurroundings heated the liquid so TINT0, and Ma was negative,but as the pressure was lowered, Ma became positive andreached 22,600 in the series of experiments. When Ma wasplotted as a function of the evaporation flux, the results obtainedare shown in Fig. 6 [26,29].

If the molar density in phase-i is denoted as ni, the fluidvelocity as vi, the enthalpy as hi, then the energy transport bybulk-phase convection and the thermal conduction in phase-i, je

i

may be expressed

jie ¼ nivihi � jijTi ð4Þ

If energy transport by surface tension-driven convection isneglected, the energy transport from the liquid to the surfacemust equal the energy transport from the surface to the vapour.Also, the mass flux from the liquid to the surface must equal themass flux from the surface to the vapour, and each is equal theevaporation flux:

jev ¼ nLvL & ir ¼ nVvV & ir ð5Þ

Then since the difference in the enthalpies is insensitive tothe temperature changes along the interface that we consider,Eq. (4) may be integrated over the surface to obtain

JeνðhV � hLÞ ¼RAðjVjTV � jLjT LÞ & irdA ð6Þ

where Jev is the measured evaporation rate.The left-hand side of Eq. (6) defines the energy transport

required to evaporate the liquid at the measured rate, and theright-hand side, the amount of energy transferred to the interface

by thermal conduction. If thermal conduction does not transferthe required amount of energy, it would suggest there must bean additional energy transport mechanism.

The possibility of an additional energy transport mecha-nism was investigated in [26] and [29] by comparing the left-hand and right-hand side of Eq. (6). In Fig. 7, the abscissa isthe energy required to evaporate the liquid at the measuredrate, and the ordinate is the energy transported to the interfaceby thermal conduction. As seen in this figure, thermal con-duction provides the energy required to evaporate the liquidin the first six experiments (EV1–EV6) at the lower evapo-ration rates. When the results in Fig. 6 are compared withthose in Fig. 7, one finds that the maximum Marangoninumber in these six experiments was 38. Thus, according tothe Pearson criterion, no surface tension driven-convectionwould have been expected in any of these six experiments,and the fact that the thermal conduction provided the energytransport required to evaporate the liquid supports the pre-diction that there was no surface-tension driven convection inthese experiments.

For the EV7, Ma was 127, and as indicated in Fig. 7, theenergy transport to the interface by thermal conduction wasslightly less than the energy transport required to evaporate theliquid at the measured rate. Thus, the Pearson criterion indicatesthat surface tension-driven convection would be present and theenergy criterion indicates that the surface tension-drivenconvection transports energy to the interface.

As the evaporation rate was increased, the deficit increasedbetween the energy transport required to evaporate the liquid atthe measured rate and the energy transport by thermal con-duction, reaching 42% at the highest evaporation rate obtained inthe series of experiments (3.38 g/m2 s). The temperature profileon the centreline at this evaporation rate is shown in Fig. 2 andparallel to the interface in Fig. 3. Thus, the energy transport bysurface tension-driven convection must transport a significantportion of the energy required to evaporate the liquid [26,29].

Page 7: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 8. Relation between the mean evaporation flux and the vapour-phasepressure when water evaporated under steady-state conditions while maintainedat the mouth of the funnel shown schematically in Fig. 1 [26].

Fig. 9. Schematic of experimental apparatus used to measure temperature closeto the water–vapour interface during steady-state water evaporation [28].

207K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

2.3. The transition to interfacial turbulence during waterevaporation

An investigation of the maximum evaporation rate that couldbe achieved with the funnel, shown schematically in Fig. 1,indicated the evaporation rate could be increased until the waterspontaneously froze at the funnel mouth. Water evaporated understeady-state conditions up to a flux rate of 4.24 g/m2 s; however,there was a transition in the interfacial flow at 3.8 g/m2 s. At thelatter flux rate, Ma had a value of 21,500, and when theevaporation flux was increased, as seen in Fig. 6, theMa numberbecame almost independent of the evaporation flux. This issignificantly different behaviour than had been seen at lowerevaporation rates, where Ma had been found to increase linearlywith the evaporation flux. It has been supposed [26,28] that thisbehaviour results from the interfacial flow becoming turbulent.We review the evidence supporting this supposition.

The relation between the pressure in the liquid phase and themean-evaporation flux is shown in Fig. 8. Note when the pressurewas reduced below 300 Pa, there was no longer a one-to-onerelation between the mean-evaporation flux and the vapour phasepressure. At 300 Pa (EV16), Ma had a value of 21,500, and areduction in pressure to 260 Pa, only increased it slightly. Theparameter that changed the most was the interfacial speed parallelto the interface: as the pressure was reduced from 300 to 260 Pa, itincreased from 0.65 to 1.15mm/s. An investigation of the viscousdissipation in the surface phase indicated it increased inproportion to the maximum-interfacial speed [29].

2.4. Interfacial temperature discontinuity during steady-stateevaporation

Earlier sections have indicated that surface tension-drivenconvection transports a significant portion of the energy required

to evaporate water. The analysis of the mechanism by which thistransport takes place is complicated by the possibility of aninterfacial temperature discontinuity. The traditional view hasbeen that the interfacial liquid and vapour temperatures are equal[31]. However, this view has been challenged as small-diameter(∼0.05 mm) thermocouples have been used to measure thetemperature close to the interface.

One of the first challenges arose from the experiments of FangandWard [28]. The apparatus they used is shown schematically inFig. 9. Both their experimental procedure and their apparatusweresimilar to those used by Ward and Duan [26]. The importantdifferences were that Fang and Ward used a glass funnel and thepositioning micrometer could only place the thermocouple on thefunnel centreline in the liquid or vapour phases, but twodifferently sized thermocouples were mounted on the positioningmicrometer. The smaller was constructed of 0.025 mm-diameter(chromel–alumel) wire and had a bead of ∼0.05 mm-diameter.The other was constructed of 0.08 mm-diameter wire, and had abead of ∼0.16 mm diameter. The objective was to measure thetemperatures in the water and vapour phases as close as possibleto the interface as water evaporated steadily, using each of the twodifferently sized thermocouples. The larger thermocouple acted toconfirm the measurement of the smaller thermocouple in thevapour phase.

A series of experiments was run with different pressures inthe vapour phase. The lowest pressure was 195 Pa and theevaporation flux at this pressure was 0.5386 g/m2 s [28]. Thetemperature profiles measured at this pressure are shown on twodistance-scales: in Fig. 10 the position-scale division is 1 mm,but in Fig. 11, it is 0.025 mm. The estimated mean-free-path

Page 8: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 11. Temperatures measured close to the water–vapour interface for theexperiment conducted with a vapour-phase pressure of 195 Pa. The generaltemperature profile is shown in Fig. 10 [28].

208 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

(MFP) in the vapour, for the experiment shown in these figures,was also 0.025 mm. Thus, the thermocouple-bead radius isequal to the MFP. This allows the centre of the thermocouplebead to be brought approximately 1 MFP from the interface.

As may be seen in Fig. 11, the smaller thermocouple could beused tomeasure the temperature in the liquid phase at a distance of∼9 bead-radii below the liquid–vapour interface. At this positionthere was essentially no distortion of the liquid–vapour interfacedue to the thermocouple. During the experiment, the open squaresin this figure are the first measurement with the ∼0.025 mm-radius thermocouple bead, and the crosses are the repeated mea-surement with this thermocouple. If one assumes, the temperaturewas uniform near the interface, as subsequent studies indicated(see Fig. 2), then the difference in the interfacial liquid tem-perature and the vapour temperature one MFP from the interfacewas 8.0 °C. However, Fang and Ward [28] solved the heatconduction equation subject to a boundary condition obtainedfrom Eq. (6). When this analytical solution was used to obtain theinterfacial temperatures in the liquid and vapour phases, itindicated the temperature discontinuity to be 7.8 °C.

A temperature discontinuity is not only a characteristic of waterevaporation. The temperature near the interface in both the liquidand vapour phases of methylcyclohexane and of octane weremeasured while each evaporated under steady-state conditions inthe same apparatus as that used to study water evaporation [32].It was found that the interfacial vapour temperaturewas greater thanthat in the liquid, and that as the evaporation rate was increased thetemperature discontinuity increased as well. The interfacial tem-perature discontinuity of water also increased as the evaporationflux was increased [33]. The maximum temperature discontinuityobserved for methylcyclohexane was 4.9 °C, and for octane, it was5.7 °C. Thus, the trends were the same for methylcyclohexane andoctane as for water, but the magnitude of the temperature discon-tinuities was smaller for the hydrocarbons.

We note that a study of the temperature profile near the interfaceof water during droplet evaporation [34] and during steady-statecondensation [35] indicated there was an interfacial temperature

Fig. 10. Centreline temperature of water as it evaporated under steady-stateconditions while maintained at the mouth of a glass funnel [28]. The tem-perature profile was measured repeatedly with two differently sized ther-mocouples (thermocouple beads: ∼0.16 mm and 0.05 mm-diameter) while thepressure in the vapour phase was 195 Pa and water evaporated steadily at a rateof 0.5386 g/m2 s. Near the interface the temperature could only be measuredwith the smaller diameter thermocouple bead.

discontinuity during each of those phase-change processes as well.This suggests that an interfacial temperature discontinuity is acharacteristic of phase-change processes in general. But the issue ismore complex than simply the flux fromone phase to the other. Forexample, in thewater evaporation experiments, conductedwith thestainless-steel funnel shown schematically in Fig. 1, when theevaporation flux was 1 g/m2 s the temperature discontinuity was2.16 °C. This evaporation flux is almost twice that of theevaporation flux from the (borosilicate) glass funnel and thetemperature discontinuity in the evaporation experiment using theglass funnel was 7.8°C. This comparison of the results obtainedfrom the funnels constructed from different materials indicates adependence of the temperature discontinuity on the properties ofthe material holding the water during the evaporation process, andnot just on the evaporation flux. But for any one funnel, providedthe interface curvature is kept constant, the interfacial temperaturediscontinuity increases with the evaporation flux [33]. The role ofinterface curvature has received some attention: Popov et al. [27]compared the temperature discontinuity during steady-state waterevaporation from the mouth of a stainless-steel channel. Theyfound a larger temperature discontinuity at a smaller evaporationflux by reducing the curvature of the interface. An analytical studyby Bond and Estrup [36], based on classical kinetic theory,suggested this is the effect interfacial curvature should have on theinterfacial temperature discontinuity.

The existence of an interfacial temperature discontinuity dur-ing both evaporation and condensation is supported by calcula-tions based on statistical rate theory. This theoretical approach tointerfacial kinetics [33,37–39] allows the pressure in the vapourphase to be predicted from the measurements of the interfacialtemperatures, the evaporation flux, and the curvature of theliquid–vapour interface. Importantly, for water there were nofitting constants in the predictions. For both evaporation andcondensation of water, the predicted pressure in the vapour phasewas not distinguishable from the measurement [33,34,30,40–42].The results reported byWard and Fang [33] are shown in Fig. 12.

Notice that all of the predictions lie within the error bars(±13 Pa) of themeasurements. One of the inputs to the calculationwas the measured temperature discontinuities. Although therewas no measured disagreement, it cannot be taken as conclusive

Page 9: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 12. Comparison of the predicted and measured vapour-phase pressure in theseries of water evaporation experiments. No fitting constants were necessary tomake the prediction.

209K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

that there was an interfacial temperature discontinuity in each ofthe experiments, because of the predictions are so sensitive to thepressure [33]. Within the error bars of the pressure measurement,even the direction of the evaporation could be reversed [40]. Theconclusion must be that the statistical rate theory calculations areconsistent with an interfacial temperature discontinuity at theliquid–vapour interface during steady state evaporation.

3. Interfacial driven flows during evaporation of non-polarliquids

Capillary driven fluid flows in non-polar liquids have attractedincreasing attention of the scientific community for the lastcentury. These effects and induced flows can be of paramountimportance in a wide range of biological and industrial areas. In arecent review Schatz et al. [14] emphasized on the distinctionbetween two categories of capillary driven flows. The first wasdescribed as Marangoni-convection, whereby the imposedtemperature gradient is perpendicular to the interface. In thisconfiguration the basic state is static with conduction temperaturedistribution; motion ensues with the onset of instability when thetemperature gradient exceeds some threshold (critical). Thesecond category was described as thermocapillary-convection,whereby the external imposed temperature gradient is primarilyparallel to the interface. In this casemotion occurs for any value ofthe temperature gradient. It is worth noting that in this secondcase, exerted capillary stress has to overcome viscous resistance inorder to have onset of motion.

Marangoni convection can be macro- or microscopic, adivision determined by the nature of the onset of the convection.This difference is important because it can affect how mixingtakes place in a container. Macroscopic convection will lead tomixing throughout the container while microscopic convectionwill only lead to mixing near the interface. Thus thisphenomenon has applications in the chemical industry, forexample in the design of mass transfer equipment and chemicalreactors where the circulation of fluid caused by solutal-Marangoni convection (i.e. depending on concentrations in the

mixture) would have beneficial effects. Berg [43] describes howthese two types of capillary flows arise:

1. Microscopic convection is brought about by self-amplifica-tion of small disturbances, i.e. the instability discussed in thework by both Pearson [7] and Sternling and Scriven [15].Take for example the case where a solute that lowers surfacetension is being transferred upwards across a liquid–vapourinterface. A small disturbance at the interface can bringsolute-rich liquid to the site of the disturbance, thus reducingthe surface tension at that point and setting up forces thatpromote further disturbances until convection takes hold.

2. Macroscopic convection results from any inherent macro-scopic asymmetry in the system which can lead toMarangoniconvection (e.g. wax flow in a candle pit due to heat source).For the microscopic instability-induced Marangoni effect tooccur, the systemmust be unstable, but whether a disturbanceactually starts to grow from the instability depends on anumber of factors, including the wavelength of the dis-turbance. The Marangoni number, Ma, is a critical numberused to establish whether Marangoni convection will occur.

It is also recognised that surface tension becomes thedominant effect in all those cases involving an interface whenthe characteristic dimension drops below the capillary lengthðlc ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffir=qgÞp

. In such small configurations, experimentaldifficulties in measuring parameters such as temperature,pressure and velocity are still an important issue.

Owing to the fact that Marangoni convection circulates thefluid in a system it can also affect mass and heat transferprocesses. The circulation increases the transport of material toand from the liquid–vapour interface, thus increasing the masstransfer coefficient. Furthermore, the same circulation of fluidincreases heat transfer rates through the system. Khrustalev andFaghri [44] investigated thermocapillary convection usinggrooved plates heated on the bottom. They set about establish-ing the influence of the liquid and vapour flows resulting fromthis effect on heat transfer from the meniscus formed by theliquid in a groove. They found that the convection currentsformed in the liquid just below the meniscus could increase theheat transfer coefficient by up to 30%. This is presumably due tothe assistance they give in bringing liquid from the bulk regionright up to the interface where the heat transfer occurs, and thenremoving it again. They also found that if there are very largetemperature differences between the wall and the vapoursaturation temperature, then recirculation of the vapour couldappear above the meniscus in the region where the meniscusmeets the container wall.

Progress on better understanding these phenomena ishampered thus far by the access to local measurements atvery small scales. The experimental techniques used toinvestigate thermocapillary flows range from Schlieren tech-nique, interferometry, holographic interferometry and veloci-metry techniques. The access to interfacial properties, which arecrucial in understanding the above phenomena, is limited by theaccuracy of the adopted techniques. Things become even morecomplicated when the interface undergoes heat and mass

Page 10: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 14. Three-dimensional plot of evaporation rate versus microwave powerand deviation from equilibrium pressure.

210 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

exchanges as in the case of evaporation or condensation. Beingprimarily driven by variations in local interfacial properties, it iscrucial to have access to precise measurements of theseproperties. For instance, till recently the sizes of availablethermocouples did not allow precise measurement of interfacialtemperatures. Many attempts have been made to accuratelymeasure interfacial temperature during the evaporation ofliquids. Despite extensive research undertaken on the subject,the present understanding of thermocapillary convection in thepresence of evaporation is far from being complete.

Shieh et al. [45] studied the case of thermocapillary-buoyancy driven convection in an insulated rectangular tankwith free surface in the middle of which a heat source is placed.The situation involved liquid undergoing evaporation at the freeupper surface. Shieh [45] clearly showed that there is significantenhancement of convection by the heat sink due to evaporationthat produces a more vigorous and efficient mixing. Shieh alsodemonstrated that Marangoni convection drives the liquid fromhot to cold regions of the flow diminishing the temperaturegradients along the interface helping in stabilizing the unstabletemperature stratification produced by evaporation. It is worthmentioning that local horizontal temperature gradients opposedto the global imposed temperature gradient between the endwalls (Villers and Plattern [46]) and vertical temperaturestratification in gravitational environment are dangeroussources of instabilities for liquids with Pr larger than one(strong coupling between thermal and momentum fields).

Recent investigations confirmed the idea that the evaporationprocess through the latent heat can drive thermocapillaryconvection without external heat supply. Mancini [47] usedhexametildisiloxane to investigate convective patterns induced bylatent heat of evaporation (Fig. 13). In the experiments performedby Mancini no external heating or cooling is applied. It isdemonstrated that when latent heat associated with evaporation islarge enough, the heat flow through the free surface of the fluidcan generate temperature gradients that can destabilize theconductive motionless state and give rise to convective motion.A typical sequence of ordered patterns appears similar to aBénard–Marangoni convection heated from below. It is foundthat the evaporation process generates a vertical temperaturegradient leading to the observed convective patterns.

Gillon [48] experimentally investigated instabilities in evap-orating liquids under low pressure and/or microwave irradiation(that will be referred to in the following as the constraint). The

Fig. 13. Sequence of capillary driven patterns durin

evaporating meniscus of ethyl alcohol under controlled pressureand/or microwave radiation has been experimentally investigated(see Fig. 14). The meniscus is maintained at the same level by amercury piston system and the evaporation rate ismeasured by thelevel of mercury into a tank. The surface temperature is measuredby IR thermography technique and the fluid motion (near theinterface) is recorded by a video camera. The evaporation rate as afunction of the applied constraint (that is pressure in the firstexperiment of the study) shows three distinct regions corre-sponding to three different regimes. The first region (low con-straint) corresponds to the purely conductive mechanism; thesecond region is the convection region in the liquid bulk phase;the third region corresponds to the boundary layer's instabilityabove the buoyancy driven convection cells.

Gillon [48] used latex particles to visualize the motion in thebulk liquid phase. At low constraints the purely conductiveregime leads to the evaporation rate's linear domain. Increasingthe constraint a motion exhibiting axi-symmetric and stationarycells in the bulk liquid phase is set; IR thermography showed thatthe roll centre is cold. As the constraint is further increased, theconvection cells become unsteady and increase in number andtheir size becomes smaller. In the transition region between thesecond and the third regions of the evaporation rate versusimposed pressure (constraint) curve; the convection cells slowdown. The authors report a decrease in the evaporation rate afterthe onset of the second instability. The decrease in evaporationrate after the second regime is attributed to competition between

g the evaporation of hexametildisiloxane, [47].

Page 11: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 15. Evaporation of Ethyl alcohol versus low pressure. Three modes ofevaporation revealed.

211K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

endothermal surface evaporation process and the thermal flux (socalled Smith–Topley effect). Increasing further the constraint,stronger new convection regime occurs, but now at a smaller scalein the boundary layer above the buoyancy cells; this effect iscalled the Marangoni instability (Fig. 15). For large values of theconstraint there is a noticeable depression of the cell's centeraccompanied by the propagation of thermals in the liquid phase

Fig. 16. Marangoni cells and vapour recoil instability (hot spots), (a)–(b) cylindrical tof a hot spot, square cavity, [48].

below the growing depression. At some point the liquid issuddenly ejected from the interface and an important increase ofthe particles' velocity has been noticed; this effect is called thethermal boundary layer differential vapour recoil instability.

The authors [48] also reported observations on theevaporation of ethanol under microwave irradiation (Fig. 16).The region of purely conductive regime is very small. Nostationary convective regime has been observed; this is due tothe autocatalytic effect: the temperature's fluctuations inducefluctuations of the dielectric loss constant and thus increment ofthe local temperature fluctuation. Instead two unsteady regimesare observed. The first unsteady regime takes place in a narrowinterval of applied constraints and is characterized by a smallslope; the particles motion is essentially axi-symmetric. Thesecond unsteady regime takes place for a broader interval ofapplied constraint. Very large surface temperature oscillationsare observed with IR thermography.

It must be pointed out as for low values of pressure; thesurface temperature decreases logarithmically as the constraint(pressure) increases; whereas for pure microwave constraint thesurface temperature increases with the constraint. In the latter

ube, (c)–(d) surface deformation with Marangoni cells, (e)–(f)–(g) amplification

Page 12: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

212 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

case the evaporation process is not fast enough to carry out theheat produced, that is used to heat the system. At very largemicrowave constraint, a new change of the evaporation rateversus the applied constraint curve is observed; this is the vapourrecoil instability regime taking place without passing throughthe Marangoni instability regime (Fig. 17). Kavehpour [49]investigated the spreading of volatile liquids on solid substrates.A non-invasive technique using a laser confocal microscope hasbeen used to investigate the evolution of viscous micro-dropletson solid surfaces. The Ohnesorge number, defined as the ratio ofthe viscous forces to the surface tension forces, and the intrinsiccapillary length has been introduced as a parameter to detect thespreading regime and which forces act in it. Hydrodynamicinstabilities can occur when there are local variations in thesurface tension, which is one of the parameters that control thespreading of drops. Marangoni instabilities are the result ofthermal variation perpendicular to the fluid interface and whenthese instabilities arise for an applied temperature field along theinterface, they are usually referred to as hydrodynamicinstabilities. Thermocapillary flow arises for differential evap-oration along the drop interface. The evaporation rate is strongernear the contact line and thus a gradient of temperature (due tothe larger cooling effect near the triple line). A dimensionlessinterfacial thermal resistance is introduced (this number isessentially the inverse of the Biot number) and by linear stabilityanalysis it is shown that there exists a critical value of thisnumber below which there is no instability. It is also shown thatas the drops spread and their height decreases, the amplitude ofthe oscillations increases and the frequency decreases.

It is shown that dimensionless interfacial thermal resistanceis inversely proportional to the film thickness and the latent heatof vaporization. When this number becomes larger than thecritical value, instabilities develop. A non-volatile liquid iscompared with a volatile one, and the results show that at thebeginning the spreading mechanism is similar for the two cases;when the drops spread out, the heat and mass transfer due toevaporation starts to have an effect and the spreading behaviourof the volatile liquid differs substantially from the non-volatileone with clear oscillations travelling from the edge to the centreof the drop due to Marangoni instabilities. Because these wavespropagate towards the hot region of the drop (apex), it can beconcluded that they are indeed hydrothermal in nature as theone described by Riley [50] and Smith and Davis [51,52]. The

Fig. 17. Evaporation of ethyl alcohol versus microwave power.

mechanism for the onset of these instabilities is explained by theauthors as: evaporation of the volatile material from the fluidfilm results in a loss of thermal energy from the fluid droplet anda consequent cooling in the droplet. Kavehpour et al. [49]emphasize that this is “a separate class of Marangoni flowsarising from evaporation. Thermal energy is removed from theliquid resulting in local changes in the temperature and thusalso in surface tension”.

It is clearly demonstrated that evaporation, through the latentheat can drive thermocapillary instabilities of evaporatingliquids. Unlike precedent investigations, the interfacial gradi-ents are self-generated by the evaporation process (notimposed). The increase of the evaporation rate is found tolead to transition through different regimes and instabilities.

Another mechanism responsible for convective motionduring evaporation is related to discontinuity in the velocityand in the rate of momentum transport at the interface. Theresult is a normal force, called the vapour recoil force, at theinterface acting in the direction of the denser fluid (Hickman[53], DeBroy [54]). When the system is at atmospheric pressurethis force is insignificant. However, under vacuum conditions itbecomes much more important, in particular for liquids that areevaporating at pressure below approximately 100 Pa. At aliquid–vapour interface the uneven temperatures lead to varyingevaporative fluxes from the interface and hence non-uniformvapour recoil forces. Therefore, any disturbance that increasesthe local surface temperature increases the recoil force andeventually becomes large enough to produce a depression on theinterface. This has slanted walls, allowing the evaporatingvapour to shear the liquid surface and drag hot liquid up to thesurface to the point where the surface temperature was alreadyelevated, thus further amplifying the disturbance. It is found thatthe potential for instabilities by differential vapour recoilincreases as the evaporation rate is increased and as the systempressure is decreased. Indeed, analysis has been carried out intothis effect which verifies its mechanism. It predicts thathorizontal evaporating interfaces become increasingly moreunstable as the evaporative flux and vapour shear forces areincreased, reducing the surface tension and making the interfacemore easily deformed. Palmer [55] discussed the importance ofvapour recoil during evaporation of a liquid under vacuum, inparticular noting that the interfacial convection it brings aboutcan greatly enhance the transport of heat and mass at theinterface. Many investigations (Miller [56], Palmer [57] andHiguera [58]) have looked in the stability of liquids undergoingintense evaporation. Experiments performed on superheatedwater exposed to very low pressures, show an instabilitycharacterized by the break up of the liquid surface and theejection of droplets. The evaporative flux in this condition isfound to be enhanced many orders of magnitude greater than theordinary evaporation. Grolmes and Fauske [59] experimentallyinvestigated the stability of a water liquid–vapour interfacesubjected to rapid depressurization (as low as 0.004 atm). Theyobserved a violent flashing of the interface and entrained liquidwith vapour. The hydrodynamic instability of rapidly evapo-rating liquids has been modelled by Miller [56]. The formerreported that an evaporating interface can be unstable due to

Page 13: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

213K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

large density difference between phases, with the stabilityanalysis showing the perturbation growth rate to be positivedepending on the surface tension, density difference and heattransfer at the interface. The latter adopted a linear stabilityanalysis to introduce a differential vapour recoil mechanism, toexplain the instability of the interface. It is worth noting that thetimescale of surface tension driven convection in liquidsundergoing evaporation is much slower than the recoil instability.

4. Interfacial conditions and thermocapillary flows inorganic liquids

4.1. Spontaneously evaporating non-polar liquids in capillarytubes

The following work studies the spontaneous evaporation ofvolatile organic liquids. The geometry investigated is smallcapillary tubes, where liquids evaporate from a meniscus. Theprincipal aim is to reveal interfacial effects and fluid flowassociated with the evaporation process. Evaporation of volatileliquids (ethanol, methanol, acetone and n-pentane) in smallcapillary tubes with internal diameters ranging from 200 to

Fig. 18. Schematic diagram o

900 μm was studied, [60–65]. The evaporation is found to driveconvection in the liquid adjacent to the interface. Convection inthe liquid phase of evaporating menisci formed in these capillarytubes is observed, studied and characterized. The evaporation ratewas monitored following the receding interface of the meniscus.Through the use of three different experimental techniques(micro-Particle Image Velocimetry, Thermochromic LiquidCrystals and Infra Red thermography), information on interfacialtemperature and flow patterns in the liquid phase were gathered.All liquids used were pure. Because the investigated phenomenaare very sensitive to contamination the tubes were cleaned usingan ultrasound bath with de-ionized water and dried in an oven.

The experimental setup used for the PIV measurements issketched in Fig. 18; it comprises a Research microscope coupledwith a high speed camera capable of 1000 full frames per secondat 512×512pixels and up to 30,000 frames per second at lowerresolution. The camera is connected to a PC on which recordedsequences are subsequently analysed using appropriate PIVsoftware. The experimental setup described above lies on anoptical low-vibration table to minimize disturbances (Fig. 18).Particles have been used as tracers in the liquid phase to visualizeand characterize the convection motion. The tracers were nylon

f the experimental setup.

Page 14: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 20. Evaporation flux as a function of tube size for various liquids.

214 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

spherical particles with mean diameter of 20 μm. This experimenthas been repeated using borosilicate glass particles with 3–26 μmdiameters (depending on the capillary size). All experiments havebeen performed in a room with air conditioning to maintainconstant ambient temperature. Humidity and pressure in the roomare monitored using a meteorological station.

The IR camera has been used to map the temperature on themeniscus interface as well as on the wall of the capillary tubes.The IR camera used in this study has a thermal sensitivity of20 mK and an accuracy of 1% (see Fig. 18). The system canacquire images at high speed (up to 750 Hz). The images acquiredare transferred to a dedicated PC with specialised software toextract temperature data. PIV and IR measurements have beenconducted simultaneously during most experiments. TLC mea-surements were conducted in a totally dark environment. Theexperimental setup comprises a 3CCD colour camera; two fibreoptic cables; the capillary tube under test and a calibrator with theformer positioned underneath the latter to enable effective use ofthe whole CCD array. Each of the three RGB components isgrabbed separately on a computer operating at 30 frames persecond; temperature data are then extracted. While TLCtechnique has been mainly used to map temperatures on thewall of the capillary tubes, IR has been used to measure bothtemperatures on the liquid–vapour interface and the tube wall.

4.2. Receding evaporating meniscus and observed flows

It is known from previous works, P.C. Wayner [66], that inthe case of a meniscus in a capillary tube the liquid evaporatesessentially in the thin film region of the meniscus in contactwith the tube wall. This has been demonstrated by manyauthors, S. DasGupta [67] and confirmed in this investigation.In what follows we briefly review the physics of an evaporatingliquid meniscus (see Fig. 19). The meniscus region close to thetube wall is usually divided into three different sub-regions(I. Kim [68], L.W. Swanson [69], Park [70]). The adsorbed layerof constant thickness where the Van der Waals forces dominateand no evaporation takes place. The micro-region which is avery small region (usually a few percent of the entire meniscuslength) where most of the evaporation takes place because film

Fig. 19. Different identified regions during the evaporation of a meniscus.

thickness is still small and the adhesion forces are balanced bythe capillary ones. The macro-region where the meniscusthickness increases, the thermal resistance of the layer increasesaccordingly and the heat transfer coefficient weakens. Usuallybigger tube sizes have a longer micro-region; the micro-regionincreases linearly with the tube radius, [70], Fig. 19.

Because of the above described wetting characteristics theevaporation along the interface of the meniscus formed insidethe capillary tube is not uniform. It is larger near the wall than inthe middle of the capillary (Moosman and Homsy [71]). Theresults obtained in this study from the measurements of theevaporation rate and flux confirm this trend. Indeed theevaporation rate, for a stationary meniscus stuck at the tubemouth, is found to be linearly proportional to the tube radius(not the tube area), Fig. 21, [71]. This suggests that most ofevaporation is concentrated near the triple contact line of themeniscus. Whilst the evaporation flux is found to increase whenreducing the tube size, Fig. 20, the overall evaporation rateincreases linearly with increasing tube size, Fig. 21. Themeasurement of the temperature profile along the tube wallusing TLCs allowed to deduce the profile of heat flux. Theresult shows a maximum in heat flux (minimum in temperature)near the contact line. The increase of the evaporation flux with

Fig. 21. Evaporation rate as a function of tube size for various liquids.

Page 15: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 23. Meniscus position inside the capillary measured from the tube mouth,results for three different liquids.

Fig. 22. (a) Marangoni convection pattern: flow visualization with the use ofordinary light. (b) PIV analysis showing the vector map with streamlines andvorticity.

215K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

decreasing tube sizes indicates an important role of the observedconvection in the enhancement of evaporation process. Indeedwhen reducing the tube size the observed convection is found tobe more intense as will be explained further.

The interfacial temperature during the evaporation of liquids inthis study is found to be non-uniform. Indeed by using the IRtechnique, interfacial temperature on the liquid side was measuredduring the evaporation of various liquids. The likely explanation forthese measurements is that the non-uniform evaporation cools theliquid phase differentially along the meniscus interface leading to adifference in interfacial temperature. The generated temperature

variations would lead to gradients in surface tension and densityalong the meniscus interface. For small tubes (below 1 mm indiameter) the surface tension effect is likely to be dominant, [72]and [73]. The observations of the fluid flow in liquid revealed aconvection roll that has a toroidal shape and looking at it from theside two sections are observed, Fig. 22a. In the spinningmovementthe particles are accelerated along the meniscus interface from thecenter to the wedge; from there they move back along the walls tothe bulk phase returning to the center of the capillary and then areaccelerated again along the capillary center towards the meniscus.PIV analysis of the observed convection reveals a detailed flowvelocity map, Fig. 22b. At smaller tubes sizes the intensity of theconvection roll increases because of a higher temperature gradientalong the meniscus interface.

The investigation demonstrated the role played by theobserved convection in bringing hot liquid from the bulk closeto the meniscus interface where evaporation is taking place.This contributes to the enhancement of heat and mass transfer.Although convection contributes to the enhancement of theaverage evaporation flux, a time-scale analysis has shown thatfor the investigated system most of the heat transfer necessaryfor sustaining evaporation comes through the walls of thecapillary.

One of the objectives of this investigation was to relate theevaporation process to the observed convection. Two types ofexperiments have been performed. In the first experiment, themeniscus has been positioned at the capillary mouth using amicro-syringe, and then was left free to recede inside the tube as itevaporates. The distance of the meniscus from the tube mouth (L)is measured as a function of time; this measurement allows theevaluation of the evaporation rate. Fig. 23 shows typicalmeasurements of the meniscus position inside the capillary tubeversus time. The presented results are for liquids of differentvolatilities. The extracted velocity of the interface shows thestages of evaporation. It is clear that the slope of the curves(corresponding to the evaporation rate) increases noticeably withsmaller tube size and with the use of more volatile liquids. The

Page 16: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 24. Temperature evolution of triple line as the meniscus moves deeper inside the tube extracted from IR measurements.

216 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

slope of each curve decreases also as the meniscus recedes insidethe tube, Fig. 23. Essentially, the slopes of the curves decrease asthe meniscus moves deeper inside the tube confirming that farfrom the tube mouth the evaporation is basically controlled by thediffusion of vapour from the interface to the tube open end.

Fig. 24 reports a typical behaviour of the temperatureevolution of the meniscus triple contact line (measured on thewall by IR) as it recedes inside the tube (ethanol in a 600 μmdiameter tube). As evaporation proceeds mass is lost and the

Fig. 25. Temperature measurement (IR) along the

meniscus moves deeper inside the tube. This movement ischaracterized by two stages. During the fist stage the largeadhesion forces keep the meniscus anchored at the capillarymouth and the meniscus changes shape. When the recedingcontact angle is reached, the meniscus detaches and movesdeeper inside the tube. The detachment event can be clearlyseen in Fig. 24 after which the triple line experiences a steepincrease in temperature. The slight temperature decrease beforedetachment is thought to be due to the thinning of the meniscus

tube wall during the evaporation of ethanol.

Page 17: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

217K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

prior to detachment, this leads to higher evaporation and no-ticeable cooling.

4.3. Temperature distribution

Recording the interfacial temperature allowed detecting thestrong evaporative cooling near the meniscus triple line.

In addition to temperature measurements on the tube wallnear the meniscus contact line, the temperature on the liquid–vapour interface has been mapped using IR camera. Thetemperature displayed on the IR image of Figs. 25, 26 and 27 isthe difference between the actual temperatures with the ambientone. Temperature measurements demonstrate the existence oftemperature gradient along the liquid–vapour interface as wellas between the bulk liquid and the interface. Indeed Fig. 25shows that a clear temperature difference between the bulkliquid and the interface exists. This observation together withinterfacial temperature measurements on Fig. 26 indicates

Fig. 26. IR results showing the interfacial temperatures for 4 different volatileliquids.

different local temperature differences between the bulk liquidand the interface. Temperature differences are obviously largertowards the edge of the meniscus.

The comparison between the different liquids showed that theinterfacial temperature is consistently lower for more volatileliquids, Fig. 26. It seems that the high evaporation rate for morevolatile liquids generates lower temperatures on the interface.

Looking closer near the meniscus wedge a clear trend isobserved. The temperature is colder near the triple contact lineof the meniscus and increases monotonously towards the centreof the meniscus, Fig. 27.

In order to relate the measurements of evaporation rates andinterfacial temperature to the observed convection, PIVanalysishas been used to quantify the flow. The simultaneous measure-ment of fluid flow velocity and interfacial temperature showedthat in some cases oscillations of both the flow and the inter-facial temperature can be present as in the case of ethanol in a600 μm diameter tube. The coupling between the oscillations inthe flow and interfacial temperature indicates the strong link andsensitivity of these phenomena (Fig. 28).

4.4. Fluid flow measurements

Simultaneously with evaporation rate measurements per-formed in the first experiment, the liquid adjacent to the interfaceis seededwith particles and the convection frequency ismeasured.The strength of the convection roll (spinning frequency) is foundto decrease as the meniscus recedes inside the tube. Theconvective roll slows down and eventually stops when themeniscus if too far from the tube mouth. The results for theevaporation rate and spinning frequency demonstrate the stronglink existing between the evaporation process and the observedconvection. The roll shape and dimensions and their evolutionhave also been monitored. As the meniscus starts to move, firstchanging its shape, the rolls are stretched in the direction of themovement. When the meniscus detaches and starts to recede, theroll shape and dimensions do not further change. As the meniscusdetaches from the capillary mouth, its profile remains unchangedalthough the driving force for evaporation changes because of thepartial vapour pressure rise; this has proved not to appreciablyinfluence the rolls shape and dimensions, but only its spinningfrequency. With the meniscus stuck at the capillary mouth, theevaporation is limited by the heat transfer from the surroundingenvironment. When the meniscus detaches from the tube mouthand starts to move inside the tube, the limiting mechanism forevaporation is the diffusion of vapour inside the tube. Results inFig. 30 show that for various capillary sizes and liquids,evaporation is reduced when the meniscus recedes inside thepore. Measurements on convection (Fig. 29) demonstrate that thespinning intensity decreases as well.

Fig. 30 shows the convection roll frequency as the meniscusrecedes inside the capillary tube. The trend shown on Fig. 30confirms the strong correlation between the evaporation rate andthe strength of the observed convection. As the meniscus recedesinside the capillary tube, the vapour generated by evaporationhas to diffuse to the mouth, the farther the meniscus is from thetube mouth the higher is the vapour pressure. This tends to

Page 18: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 27. Interfacial temperature profile near the contact line extracted from IR measurement during the evaporation of ethanol.

218 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

reduce the evaporation rate and consequently reduce theconvection spinning strength as observed on Fig. 30. It isworth mentioning that the convection is more intense for smallertube sizes as surface tension gradients are larger. The convectionfrequency is plotted as a function of the evaporative flux, Fig. 31.The results show a very clear correlation between the twoparameters. This demonstrates a very strong link betweenevaporation and the observed convection. In fact it is reasonableto assume that the convective pattern is a direct result of theevaporation process.

The second experiment consists of partially filling a tube andleaving both ends open. In this case two menisci are formed.The evaporation takes place at both menisci inside the capillarytube; one meniscus remains stuck at the tube mouth due to thestronger adhesion forces while the other recedes inside thecapillary. However, the evaporation at the receding meniscus isof less importance. The velocity of the receding meniscus isexactly the rate of evaporation taking place in the meniscusstuck at the capillary mouth. The liquid on the meniscus stuck atthe capillary mouth was seeded with particles to characterizeconvection. The evaporation rate and the spinning frequency ofconvection rolls have been correlated. In order to compare theaverage evaporation flux, the results on evaporation rate havebeen referred to the spherical cap area of the meniscus. Animportant result from this investigation demonstrates that whilethe evaporation rate is proportional to the tube radius, theevaporative flux increases when decreasing the tube size,Fig. 20. This result together with the observation of strongerconvection for smaller tubes suggests the role played by thethermocapillary convection in the enhancement of the evapo-ration flux. In order to point out the strong link between the

evaporation process and the observed convection, the averageevaporation flux and the tracers' spinning frequency have beencorrelated in Fig. 32 for the case of ethanol. Once more thestrong correlation between evaporation and the intensity of theconvection is demonstrated for the case of a stationarymeniscus. The result on Fig. 32 for a stationary meniscus andvarious tube sizes is similar to the one on Fig. 31 for a recedingmeniscus. A very strong link between evaporation flux andconvection intensity is demonstrated.

4.5. Imposed external heating and reversal of the flow

The second step in the study involved the investigation of theobserved convection for a meniscus in a capillary tube where atemperature gradient is applied along the wall. A capillary tubeis plunged into a pool of a volatile liquid; the liquid is allowed torise under the capillary effect, Fig. 33. A heater is placed at theedge of the capillary to provide localised heating. A temperaturegradient is established along the capillary tube wall whenswitching the heater on. The temperature gradients thusgenerated are measured using IR technique, Fig. 34.

The measurement of temperature gradient has been per-formed for different situations to allow a full understanding ofthe imposed temperature gradient, Fig. 35. Empty capillaries arefound to exhibit higher temperature gradients.When the heater isswitched on with an evaporating meniscus in the tube, lowertemperature gradients are reported. This is probably due to theheat dissipation brought by evaporation. It is worth noting thattemperature gradients are highest for a meniscus placed belowthe heater than for a heater placed in the liquid side below themeniscus.

Page 19: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 28. Coupling between oscillations of the flow, interface position (a) andinterfacial temperature (b).

Fig. 30. Convection roll frequency within the meniscus as it recedes inside thecapillary, ethanol in various tube sizes.

219K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

Visual observations together with PIV measurements allowdetermination of the influence of the temperature gradient alongthe wall on the thermocapillary convection described in previoussections. The μ-PIV investigation showed that the convective

Fig. 29. Extracted velocity of the interface, tube diameter, 900 μm.

patterns in the meniscus liquid phase can be altered andeventually inverted by applying sufficient extra heating. Thishappens when the meniscus lies in the tube below the heatingelement. No inversion has been noticed when the meniscusinterface is above the heating element. This is an importantexperimental evidence of the fact that the convection underinvestigation is indeed thermocapillary driven and the temper-ature profile along the meniscus interface is the key factor. UsingPIV measurement it is shown that without imposing atemperature gradient, the self-induced gradient due to evapora-tive cooling effect generates convection spinning from the centreof the meniscus towards the edge. When a temperature gradientis imposed the convection is inverted to spin from the edgetowards the centre of the meniscus. The analysis performeddemonstrates that the inversion in the convection pattern is dueto an inversion of the interfacial temperature gradient, Fig. 36.

The above described inversion of convection is found tooccur for a threshold in temperature gradient (heater power),Fig. 37. Very low power is found to be sufficient to reverse theflow. This indicates that temperature gradients originallyresponsible for convection are very sensitive to external heating.

Fig. 31. Convection frequency as a function of the evaporative flux, ethanolmeniscus receding in 900 microns tube.

Page 20: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 32. Experiment (2): tracer spinning frequency vs. evaporation flux forethanol and various tube sizes (900 μm, 600 μm, and 200 μm).

220 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

Reversing the flow by imposing external heating demon-strates the thermocapillary nature of the observed convectioninitiated by evaporation of organic liquids.

5. Discussion

The review about Marangoni/thermocapillary studies pre-sented in the Introduction section showed that the overwhelm-ing majority of investigations in the literature have dealt withcases where external heating, cooling or temperature gradientsare imposed. Studies of evaporation-driven thermocapillaryconvection in liquids and especially in water are scarce. Manyissues about the existence of thermocapillary convection drivensolely by the evaporation process are still debatable. Recentresults obtained from two independent investigations – one ofwater and one of organic liquids – have motivated this study.

The first study presented dealt with steady evaporation ofwater under variable pressure. The criterion for the presencethermocapillary convection proposed by Pearson [7] was found to

Fig. 33. Experimental setup fo

be consistent with the results obtained. It is suggested that theearlier studies [21–25] that had not been able to detect to thepresence of thermocapillary convection during water evaporationfailed because the measurement techniques used in those studieswere not able to measure the temperature close enough to theinterface. Certainly, measurements made with 25-μm-diameterthermocouples at a stationary but evaporating water–vapourinterface indicate the presence of a uniform-temperature layerbelow the water–vapour interface, an indication of thermocapil-lary convection. At one evaporation rate, the thickness of thislayer varied with position, but had its maximum value on thecentreline. The measured temperature profile parallel to thespherical water–vapour interface indicated the temperature wasparabolic with the maximum at the rim of the stainless-steelfunnel holding the water and a minimum at the funnel centreline(see Fig. 14). Such a temperature profile would be expected toproduce a surface tension-driven flow from the rim toward thecentreline, and the existence of such a flow was demonstratedwith a (13-μm-diameter, 52 mm long) cantilevered flow probe.By assuming the deflection of the probe resulted from the dragforce of the convective flow, it is possible to determine the flowspeed. Importantly, this speed could also be calculated from themeasured temperature gradient parallel to the interface. Theseindependent methods gave the samemagnitude to the flow speed.Thus, the temperature gradient appears to be the source of theobserved convection.

The source of the temperature gradient was indicated to bethe thermal constraints: the water was held in the stainless-steelfunnel, and the temperature at the funnel throat was maintainedat ∼3.5 °C. Since the thermal conductivity of water is muchlower than that of stainless steel, the evaporation at the funnelmouth lowered the temperature there below that at the throat.Thermal energy was transported along the funnel walls to therim of the funnel. This gave the parabolic interfacialtemperature profile seen in Fig. 3, and the gradient in surfacetension that produced the interfacial convection.

We speculate that the fluid flow coming from the circularfunnel rim converged on the centreline where it had to penetrate

r a heated capillary case.

Page 21: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 34. Temperature gradient measure with IR camera as heat is applied at the edge of the capillary.

221K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

the bulk liquid, but since the interfacial liquid was lighter (butcolder) than the bulk, its penetration was resisted by buoyancy;thus the return flow would be forced toward the interface. It wasthe interaction of these two streams that produced the mixingrequired to give the uniform-temperature layer.

The results from this study can be compared, at least in aqualitative way, to the previous one. The correlation between theevaporative flux and theMarangoni number can be compared to thecorrelation of the evaporative flux and the spinning frequency ofconvection. The trends from both studies are indeed in agreement.Larger evaporation fluxes exhibit greaterMarangoni numbers fromthe first study and faster convection from the second study.

The second study presented in this paper is an investigationof evaporative driven convection taking place in capillary tubesduring the spontaneous evaporation of organic liquids. Theinvestigation studied thermocapillary convection driven byevaporation. During evaporation of volatile liquids from a

Fig. 35. Temperature gradients measure with IR camera as heat is applied withvarious powers.

curved meniscus formed in a capillary tube, non-uniformevaporation cools the liquid phase differentially along themeniscus, leading to a difference in interfacial temperature andthus in surface tension and density along the liquid–vapourinterface. At small sizes, the surface tension gradient is expectedto be the driving force for the thermocapillary (Marangoni)convection roll visualized in the liquid phase. When a meniscusis stuck at the capillary mouth, the evaporation is limited by theheat transfer from the surrounding environment. When themeniscus detaches from the tube mouth and starts to recedeinside, the limiting mechanism for evaporation is the diffusionof vapour from the interface to the mouth of the tube. Resultsshow that for various capillary sizes and liquids, evaporation isreduced when the meniscus recedes inside the tube. Measure-ments on convection demonstrate that the spinning intensitydecreases as well. It was possible to correlate the convectionfrequency and the evaporative flux. Because of diffusion limi-tation, the vapour pressure near the interface increases as theinterface recedes deeper inside the tube. This situation is ana-logous to an experiment where the vapour pressure is increased(or decreased). The second important conclusion from the firststudy shows the effect of imposing external temperature gra-dient. Applying external heating and imposing a temperaturegradient along the tube wall, the above described convection isslowed down, accelerated or reversed, depending on thelocation and intensity of the heating. This result confirms thethermocapillary nature of the convection.

The two studies use different liquids in different geometriesunder different conditions; however the conclusions reached arevery similar. Both studies demonstrate that evaporationcan generate a convection of thermocapillary nature. The sogenerated convection increases in intensity as evaporation rate isincreased. Another important finding from these studies is thatthe evaporation can establish interfacial temperature gradientswhich dictate the strength and orientation of the observed

Page 22: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 36. PIV results showing the inversion of the convection as heat is applied to the edge of the capillary.

222 K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

thermocapillary flow. Despite the fundamental differences in theexperimental approaches and procedures between the twostudies important analogies between the two can be proposed:a receding interface of an evaporating meniscus inside a tubeshows that the generated thermocapillary convection graduallyslows down and eventually stops as it recedes. The convection canalso be restarted when the interface is pushed towards the tubemouth. Temperature measurements of the coldest region of themeniscus show that as the interface recedes inside the tube thetemperature tends to gradually increase. The temperature gradientalong the interface is thus reduced. The very same gradientincreases when the meniscus moves towards the tube mouth.

The analysis of this result shows that vapour diffusion fromthe interface to the tube mouth is the controlling mechanism of

Fig. 37. Vorticity versus heater power for methanol as

evaporation. The vapour pressure near the interface builds up asthe meniscus recedes inside the tube. This leads to a slow downof evaporation, temperature gradient and convection intensity,Fig. 38. The experimental procedure in the second studyinvolved reducing the ambient pressure and making measure-ments about evaporation rate, temperature profile and thermo-capillary convection. Reducing the pressure is found to increasethe evaporation rate as well as the temperature gradient alongthe interface leading to stronger thermocapillary convection.These conclusions from the two studies are – in a qualitativeway – very analogous.

The second study has also demonstrated the role played bythermocapillary convection in the energy balance at the inter-face. It is found that the generated convection has a major

liquid and heater in the meniscus vapour phase.

Page 23: Recent advances on thermocapillary flows and interfacial ... · Recent advances on thermocapillary flows and interfacial conditions during the evaporation of liquids Khellil Sefianea,⁎,

Fig. 38. Meniscus temperature and convection frequency as the meniscusrecedes inside capillary tube.

223K. Sefiane, C.A. Ward / Advances in Colloid and Interface Science 134–135 (2007) 201–223

contribution in bringing the necessary energy for the evapora-tion process. Thermocapillary convection driven by evaporationmust be accounted for in the energy balance at the interface.This is a major finding as previous accepted theories do notconsider but conduction and buoyancy driven convectionmechanisms.

References

[1] Tomson J. Philos Mag 1855;10:330.[2] Marangoni CGM. Ann Phys Chem (Poggendorff) 1871;143:337.[3] Van der Mensbrugghe GL. Acad De Belgique: Mem Couronnés et Mem

Des Savants étrangers; 1869. p. 34.[4] Marangoni CGM. Tipografia Fusi, Pavia; 1865.[5] Bénard H. Rev Gén Sci Pures Appl 1900;11:1261.[6] Bénard H. Ann Chem Phys 1901;23:62.[7] Pearson JRA. J Fluid Mech 1958;4:489.[8] Levich VG, Krylov VS. Annu Rev Fluid Mech 1969;1:293.[9] Levich V.G., Prentice-Hall, Englewood Cliffs, N.J. 1962.[10] Kenning DBR. Appl Mech Rev 1968;21:1101.[11] Davis SH. Annu Rev Fluid Mech 1987;19:403.[12] Normand C, Pomeau Y, Velarde MG. Rev Mod Phys 1977;49:581.[13] Velarde MG, Castillo JL. In: Zierep J, Oertel Jr H, editors. Convective

transport and instability phenomena; 1982. Chapter 2.1.1.[14] Schatz MF, Neitzel GP. Annu Rev Fluid Mech 2001;33:93.[15] Sternling CV, Scriven LE. AIChE J December 1959;5:514.[16] Scriven LE, Sternling CV. Nature 1960:187.[17] Velarde MG. In: Sorensen TS, editor. Lecture Notes in Physics, vol. 105.

Springer-Verlag; 1978. p. 260.[18] Lucassen J. Part I. Theory. Trans Faraday Soc 1968;64:2221.[19] Lucassen J. Trans Faraday Soc 1968;64:2230.[20] Gouda JH, Joos P. Chem Eng Sci 1975;30:521.[21] Barnes GT, Hunter DS. J Colloid Interface Sci 1981;88:473.[22] Barnes GT, Fehler AI. J Colloid Interface Sci 1980;75:584.[23] Schreiber D, Cammenga KH. Ber Bunsenges Phys Chem 1981;85:909.

[24] Cammenga HK, Schrieber D, Rudolph B-E. J Colloid Interface Sci1983;92:181.

[25] Cammenga HK, Schreiber D, Barnes GT, Hunter DS. J Colloid InterfaceSci 1984;98:585.

[26] Ward CA, Duan F. Phys Rev E 2004;69:056308.[27] Popov S, Melling A, Durst F, Ward CA. Int J Heat Mass Transfer

2005;48:2299.[28] Fang G, Ward CA. Phys Rev E 1999;59:417.[29] Duan F, Ward CA. Phys Rev E 2005;72:056302.[30] Ward CA, Stanga D. Phys Rev E 2001;64:051509.[31] Landau L.D., Lisfshitz E.M., Fluid Mechanics, (Pergamon London) P.52.[32] Fang F, Ward CA. Phys Rev E 1999;59:441.[33] Ward CA, Fang F. Phys Rev E 1999;59:429.[34] McGaughey AJ, Ward CA. J Appl Phys 2002;91:6406.[35] Ward CA, Stanga D. Phys Rev E 2001;64:051509.[36] Bond M, Struchtrup H. Phys Rev E 2004;70:061605.[37] Ward CA. J Chem Phys 1983;79:5605.[38] Ward CA, Findlay RD, Rizk M. J Chem Phys 1982;76:5599.[39] Elliot JAW, Ward CA. J Chem Phys 1997;106:5667.[40] Rahimi P, Ward CA. Int J Heat Mass Transfer 2004;47:877.[41] Rahimi P, Ward CA. Int J Thermodyn 2005;8:1.[42] Ward CA. J Non-Equilib Thermodyn 2002;27:289.[43] Berg J. Can Metall Q 1982;21:121.[44] Khrustalev D, Faghri A. J Heat Transfer August 1996;118:725.[45] Shieh C.Y., PhD Dissertation Thesis, University of Michigan, (1985).[46] Villers D, Plattern JK. J Fluid Mech 1992;234:487.[47] Mancini H, Maza D. Europhys Lett 2004;66:812.[48] Gillon, Courville P, Steinchen A, Bertrand G, Lallemant M. PCH

Physicochem Hydrodyn 1988;10:149.[49] Kavehpour P, Ovryn B, McKinley GH. J Colloid Surf A Physiochem Eng

Asp 2002;206:409.[50] Riley R.J., PhDDissertation Thesis, Georgia Institute of Technology, (1996).[51] Smith MK, Davis SH. J Fluid Mech 1983;132:119.[52] Smith MK, Davis SH. J Fluid Mech 1983;132:145.[53] Hickman K. Ind Eng Chem 1952;44:1892.[54] DebRoy T, David SA. Rev Mod Phys 1995;67:112.[55] Palmer HJ. J Fluid Mech 1976;75:487.[56] Miller CA. AIChE J 1973;19:909.[57] Palmer HJ, Maheshri JC. Int J Heat Mass Transfer 1981;4:117.[58] Higuera FJ. Phys Fluids 1987;30:679.[59] Grolmes MA, Fauske HK. Proceeding of the fifth international heat

transfer conference, Tokyo, vol. 4; 1974. p. 30.[60] Buffone C, Sefiane K. Int J Multiph Flow 2004;30:1071.[61] Buffone C, Sefiane K. Exp Therm Fluid Sci 2004;29:65.[62] Buffone C, Sefiane K. Exp Fluids 2005;39:99.[63] Buffone C, Sefiane K, Christy J. Appl Therm Eng 2004;24:1157.[64] Buffone C, Sefiane K. Phys Rev E 2005;71:1.[65] Buffone C, Sefiane K, Christy J. Phys Fluids 2005;17:1.[66] Wayner P, Kao K, Lacroix L. Int J Heat Mass Transfer 1976;19:487.[67] DasGupta S, Schonberg JA, Wayner P. J Heat Transfer 1993;115:201.[68] Kim I, Wayner P. J Thermophys Heat Transf 1996;10:320.[69] Swanson LW, Herdt G. J Heat Transfer 1992;114:434.[70] Park K, Lee K. Int J Heat Mass Transfer 2003;46:4587.[71] Moosman S, Homsy G. J Colloid Interface Sci 1980;73:212.[72] Molenkamp T., Marangoni Convection, Mass Transfer and Microgravity,

University of Groningen PhD Thesis (1998).[73] Szymczyk J. Can J Chem Eng 1991;69:1271.