Você está na página 1de 13

SYNOPSIS

An Analytical Formulation of Evaporation driven fluid Flow and


associated Transport properties in a Thin, Evaporating Droplet
Synopsis of thesis for Partial Fulfillment of the requirements for the Award of the Degree
of
Master of Technology
by
Arijit Patra
Roll No.: 11ME32001

Under the supervision of


Professor P.K.Das
Department of Mechanical Engineering
Indian Institute of Technology, Kharagpur
May, 2016.

ABSTRACT
We present a detailed theoretical study of the evaporation of a sessile droplet on a flat, impermeable
solid substrate, which is completely wettable. In a situation where evaporation is significant, the drop
spreading can be neglected, and instead the drop radius goes to zero in finite time. The evaporation rate
is limited by the molecular diffusion of vapour into the surrounding atmosphere. Essentially, the
evaporation rates and the net evaporative flux are calculated by means of obtaining the solution to the
Laplace equation to derive a concentration field. The net flux so obtained is utilised for analysing other
pertinent droplet characteristics and their variation during the evaporation process. Finally we present
variations of the evaporation flux with contact angle and radius and that of the contact angle with time
and also for the temperature gradient with contact angle and radius. Following the analysis of the
domain outside the droplet, where diffusion limited evaporation occurs; we focus on the convective
flow field inside the droplet. For this, the momentum and the continuity equations inside the spherical
cap droplet become relevant. On consideration of the droplets of very dilute suspensions, with
neutrally buoyant particles being the solutal components, an attempt is made to enhance the solutions
of the Stokes formulation by introducing the Thermal-Marangoni convection effects through the
temperature induced surface tension gradients on the droplet surface. This coupling enabled the
computation of the unsteady velocity field in the droplet, which by extension may be used as an
indicator of various other parameters like the deposition rate and the critical temperature filed required
to suppress the coffee-stain effect.

INTRODUCTION
The general process of evaporation is a basic physical phenomenon that is a part of everyday
observation. The standard definition of Evaporation as the conversion of the liquid phase to vapour
phase is simplistic at best, for it leaves out various other intricacies that govern the complex interplay
of various factors that evaporation entails. One of the pertinent issues that have aroused much
scientific curiosity in recent years is the evaporation of droplets, both of pure liquids and of
suspensions. The second case is particularly of interest because of the immense diversity of practical
applications that emanate from the associated dynamics. Although of common occurrence as the
deposit left behind by a drying droplet of coffee, which compelled Deegan et al to name it as the
Coffee-Stain Effect, the physics associated is complex and rich with potential novel implications for
inkjet printing, micro-manufacturing and myriad other areas.
The fundamental laws governing evaporation essentially put it across as a process rate-limited by two
phenomena- diffusion induced concentration gradients and the actual phase transfer from the liquid to
the vapour phase by the absorption of latent heat from the bulk. The most common approaches are a
droplet on a rough, heterogeneous surface with a pinned contact line, and one with an extended thin
film a few molecular layers thick (with due influence of the van der Waals forces manifested
significantly in such domains). We would go on to consider the latter case but it is worthwhile to first
get a broad picture of the former as well. Essentially evaporation occurs from the surface of the
droplet, and this evaporation flux is directly proportional to the effective surface area (the ratio of the
curved surface area to the volume covered) as per Ficks first law of diffusion. This creates an
interesting situation when we consider local flux across the surface at different radii- since the volume
enclosed by a chosen differential area on the curved surface would vary with the radii while the surface
area stayed constant. This leads to an increasing effective surface area as one moves away from the
centre and consequently a greater evaporation rate towards the edge. This imbalance in the evaporation
rates leads to a mass transfer inside the droplet. Further, the conditions of varying surface temperatures
leads to dynamic alterations of the surface tension, which creates a surface stress called the thermal
Marangoni stresses (a similar variation of surface tension can also be caused by the variation of
concentrations, and that is called solutal Marangoni flow, which is relatively less pronounced for dilute
solutions). This creates a convective flow on its own accord and needs to be accounted for in the mass
flux and deposition rate computation.
In the present work, we attempt to understand the evaporation of a droplet of a pure solvent, with a
focus on coupling the viscous flow problem in a thin droplet to the concentration-limited evaporation
flux using the nonlocal condition and a moving contact line. This is done as a precursor to lay down
the dynamics of droplets under a similar regime, with the novelty being the incorporation of the
Marangoni stresses in the overall velocity fields, and the deposition rate.

PROBLEM OBJECTIVE
The present aim of the problem is to analyse the evaporation for a thin and slender droplet. The
presence of a non-uniform evaporative flux gradient on the curved surface induces a convective fluid
flow from the bulk to the periphery. The combined dynamics may be mathematically analysed by
means of the Laplace equation and the Stokes flow formulation. The evaporation itself causes the
removal of latent heat from the bulk of the fluid (which reduces the internal energy of the droplet).
This creates a surface temperature gradient (which also turns out to be amenable to a Laplace
formulation) At the same time, the diffusion of the vapour molecules from the surface to the ambient is
slower than the evaporation, and hence is the rate limiting step, and this essentially creates a regime of
diffusion limited evaporation where the concentration distribution can also be worked out by using a
Laplace equation. Overall, the solution also accounts for the thermal Marangoni flows by the
implementation of a zero surface shear condition to derive the Marangoni stresses due to temperature
driven variations in surface tension.
METHODOLOGY
Essentially, every droplet on a substrate can be treated with varying degrees of complexity in the
modelling. The most common approaches are a droplet on a rough, heterogeneous surface with a
pinned contact line, and one with an extended thin film (precursor film) a few molecular layers thick
(with due influence of the van der Waals forces manifested significantly in such domains). Essentially
evaporation occurs from the surface of the droplet, and this evaporation flux is directly proportional to
the effective surface area (the ratio of the curved surface area to the volume covered) as per Ficks first
law of diffusion. This creates an interesting situation when we consider local flux across the surface at
different radii-since the volume enclosed by a chosen differential area on the curved surface would
vary with the radii while the surface area stayed constant. This leads to an increasing effective surface
area as one moves away from the centre and consequently a greater evaporation rate towards the edge.
This imbalance in the evaporation rates leads to a mass transfer inside the droplet, in a replenishment
current. Considering the volume enclosed by a given transverse cross-section at a given distance from
the centre, we can take it as a control volume and establish the mass conservation equation, which
would yield a net velocity field in the droplet. In fact this is more intuitively derived from the simple
continuity equation applied across the control volume. The dynamics incipient in the process is further
enriched when we take into account the curvature of the droplet and interpret it as a horizontal liquid
column of decreasing height in a given radial direction, and couple this intuition with the simple
Bernoulli equation, to obtain a pressure differential that mimics a capillary pressure and tries to ensure
an outward motion in the radial direction.:

In totality, the rich dynamics and the complex interplay of competing of cooperating factors leads to
the following observations in the dynamical system of an evaporating water droplet:
1. Convective motion of the fluid occurs from the centre to the periphery aided by the differential
heights, and also the variable evaporation rates over the surface.
2. Since the latent heat removal is not uniform throughout as the evaporation rates vary radially, there
exists a temperature gradient in the bulk. This causes a thermal Marangoni flow from the periphery to
the centre.
3. The existence of differential height and a varying velocity field leads to a radial pressure differential
that causes a capillary flow radially
Additionally, if the presence of solute particles are considered, rich dynamics manifested in the
phenomena of concentration driven osmotic pressure gradients, and also a solutal Marangoni flow may
be observed (in the present work, the assumption of a very dilute, ideal solution implies that the
thermal Marangoni flows are much more dominant as compared to solutal Marangoni effect which are
neglected). Presently we assume a uniform concentration field over the surface and the temperature
gradient to be zero initially, which is adjusted later on in another formulation. We also make the
requisite assumptions of no slip, incompressible flow over rough surfaces (that would be adequate for
a pinned contact line to develop), and of thin droplets where the lubrication approximations are
applicable. Additionally, we assume that the evaporation is solely limited by diffusion, and not by
energy transfer. After this part of the problem, we change the problems ambit to include the presence
of particles in a very dilute suspension. For this regime, we focus on the Stokes solution which enables
us to obtain a velocity field To extend on that, the full Navier Stokes equations are built with the stress
terms expressed through the variation of surface tension forces, and the unsteadiness accounted by
considering an average front velocity, which inherently includes a height term at a given radius and
this height is a function of time. Evaporation causes changes in temperature, which causes a surface
tension gradient creating a stress field called marangoni stress, it is balanced by a shear stress in the
system, which drives the flow in the droplet is added to the full Navier stokes equations, which helps
take into account the unsteady character. At the surface, assuming a thin drop, the radial variation of
the surface tension gives the stress.

Fig 1: The variable rates of evaporation over the surface are shown. The thinner the droplet is locally,
the greater is the local evaporation rate- essentially implying greater evaporation at edges.
MATHEMATICAL FORMULATION
We begin by deriving the model equations which describe evaporation from the drop and its
surroundings. We consider the case that evaporation is limited by diffusion. It is assumed that the
process is so slow that thermal gradients do not play a role and that the temperature T is constant
throughout the time domain considered for the evaporation.
The fundamental principle here is that we are considering the diffusion of vapour into the ambient as
our rate-limiting step of the overall process. This enables us to obtain the vapour concentration in
relation to the mass flux due to evaporation.

Fig 2: The local control volume in a transverse droplet section, to explain the notion of velocity
changes as caused due to evaporative flux imposed in the system with a variable height.

A. Evaporative flux
If fluid is confined to a closed container and equilibrium conditions are established, there will be no
net evaporation. The evaporation time of the droplet (order 0.001 s) is slow compared to the time that
it takes to build up the concentration field around the droplet (order 0.01 s). Fluid from the droplet has
to transfer from the droplet towards the air surrounding the droplet, thereby crossing the interface. This
transfer rate is characterized by a time scale of order 10-10 s. The time it would take for diffusion to
build up a profile around the droplet is of order R/D which is about 0.01 s [5]. Hence, the problem is
quasi-steady and the rate limiting step is given by the diffusion time of the vapour in air. Fick's second

law can be used to find the concentration field In the case of a macroscopic, flat interface, the mass
concentration c of vapor per unit volume of air will be the saturation concentration csat. An
evaporative flux arises by maintaining a nonequilibrium state: the concentration far from the drop is
kept at a lower value (c<csat). The transport of liquid from the drop to infinity occurs by diffusion,
where the mass flux is proportional to concentration gradients.
For any sessile droplet evaporation system, there are two major mass transfer processes involved.
Firstly, transfer of liquid molecules form liquid to vapour phase along the droplet interface involving
latent heat removal from the droplet bulk. Secondly, diffusion (natural or forced) of the vapour
molecules from the droplet interfaces to the ambient due to chemical potential gradient. In our system,
without any external agency in the ambient to carry the vapour molecules, the slowest step can safely
be considered to be the diffusion process. Thus the mass transfer process can be simplified to be an
extremely slow process with very low Reynolds number

. Thus, the process always remains

at equilibrium (quasi-steady process) and the advective transport terms can be safely neglected
compared to the diffusive terms in the ficks second law for mass transfer. So, the final governing
equation for component transport is the Laplace equation in concentration, in spherical coordinates.
Also, depending on the evaporative gradient driven heat removal from the droplet curved interface and
non uniform conductive heat flow from the isothermal solid substrate to the droplet bulk causes the
formation of a surface tension gradient on the interface as a manifestation of both surface temperature
gradient and solute concentration gradient. The only surface flow involved in the distribution of
surface temperature is surface tension driven solvent flow due to either thermal gradient
(thermocapillary flow or thermal Marangoni flow) or solute concentration gradient (solutal Marangoni
flow). For very viscous solutions, these flow velocities can be assumed to be vanishingly small in
magnitude and at the low Reynolds number limit

, advective terms in the energy equation can

be neglected to simplify it into Laplace equation in two dimensions. Also, the substrate
temperature

, thermal conductivities of solid substrate and liquid bulk can be considered to be

constant during the evaporation process and at the droplet interface, heat transfer due to conduction is
exactly equated to the rate of latent heat removal by the solvent molecules. Thus the energy equation
for this case is simplified into Laplace equation in spherical coordinates.

j Dc

(1)

Here D is the diffusion coefficient. Essentially, we have to solve the Laplace equation to enable the
derivation of the concentration field.

2c

(2)

Now the boundary conditions are to be defined. At the infinity, c=cinf is the condition, where cinf is
taken as a finite constant. Near the immediate vicinity of the drop surface, the saturation concentration

csat is the requisite condition for a flat surface and is evaluated later or explicitly provided in some
situations.
We consider the drop is thin and lubrication theory is applicable, and therefore adapt the diffusion
problem accordingly. We solve eqn(2) in the half-space directly for concentration.
It is to be noted that the evaporation volume flux obtained from the concentration is:

jev

D c
z

(3)

And the total evaporation flux is given by:


r

jtotal rjev dr

(4)

B. Stokes Flow
For highly viscous solutions droplets, flow inside the droplet is assumed to follow low Reynolds
number hydrodynamics, so temporal and advective force terms compared to the viscous terms in
Navier-Stokes equation may be ignored. These conditions simplify the steady-state Navier-Stokes
equation to produce the Stokes equations, which govern the special kind of flow, known as Stokes
flow, characteristic of highly viscous fluids in very thin drops. The convective flow of Newtonian
incompressible fluid inside the droplet bulk thus stays in equilibrium, without time dependence and
body force terms. Stokes equation, for Newtonian, incompressible fluids, can be solved using streamfunction method in both two and three dimensional axisymmetric geometries. For the case of threedimensional spherical geometry, stokes stream function formulation of governing differential equation
becomes:

where the operator

For small thin drops, equilibrium contact angle,

, remains small and the governing equations can

be simplified using lubrication approximations. Also, the flow velocity is considered to be small. This
happens when we consider the solution to be very viscous, the length-scales of the flow are very small or
the fluid flow is very slow. For fluid flow inside the droplet, the governing differential equations are
altered accordingly:
Continuity equation
Momentum equations
Vorticity equation
Where

is the vorticity. For Irrotational flow field, these simplified equations can be

expressed by stream function/velocity potential formulation which yields a biharmonic equation as the
governing differential equation. This equation can be solved to determine the stream function and
subsequently the velocity vectors inside the droplet.

As such, the average front velocity becomes directly expressible in the continuity equations, and also in
the momentum equations, which have already been simplified by the lubrication assumption. The velocity
field as evaluated from the latter is substituted into the former, after the constants of integration are
tackled with the aid of the boundary conditions. It is through this boundary condition that the implications
of the Marangoni stresses are brought into the mathematical model, by means of the zero shear stress
condition of the surface which leads to the viscous stresses and the stress due to the variation in surface
tension be equated. The resulting equation develops as a differential equation for the droplet height for a
given radial location. This is solved by the separation of variables, and the velocity is calculated from it
using the expression derived from the momentum equations (includes a term for the height).
RESULTS
We have considered a system of a droplet with a radius of 5 mm and a domain 100 times the radial
dimension so as to account for the requirements of infinite ambient. The diffusivity has been assumed at
4x10-6 for water at 298 K. As such all other parameters are set with respect to water at 298 K, with the
density explicitly set at 1000 kg/m3, the thermal conductivity at 615.4x10-3 and the specific latent heat at
2257 KJ/Kg/K. With these parameters, we set out to solve the Laplace equation and obtained the solution
by separation of variables. Further, the Stokes flow is solved for the stream function, by separation of
variables. Finally, the velocity fields are separately obtained by solving the momentum and continuity
equations, with the help of the height obtained by separation of variables in the continuity equation with a
substitution for the velocity expression. The derivative of the Laplace pressure is obtained from Popov
(2005), and is directly used in the momentum equation.
First of all the net local evaporation rate jev/is plotted with respect to the radial distance, based on
eqn(4). The radius of the droplet was taken at R=0.5 mm, and the expected singularity of the
evaporative flux was manifested at the contact line as evidenced by the sudden asymptotic rise. The
consideration of the thin precursor film, where we assumed a constant small evaporation rate, is
evident after the spike. Prior to that, the rate expectedly shows an increase as we progress towards the
edge of the droplet.

Fig 3: The variation of evaporation rate with respect to radial distance (in mm) for a droplet of
R=0.5mm. The sudden peak corresponds to r=R, and indicates the singularity at the edge of the droplet.

Additionally, it is pertinent to analyse the variations of other properties. The variation of contact angle
was indicative of interesting dynamics with respect to the total flux. It may be observed from Fig (4)
that the variations in the total evaporation for a range of contact angles is a periodic function, which is
simply because the contact angle will always be acute. The tendency to have a minima of evaporation
flux at a certain contact angle is of pivotal importance in several real-life applications. Equaally
important is the observation about the tendency to tend to infinity at angles close to 90 degrees (in a
generally observed trend of the non-linear increase of the evaporation flux with contact angle).

Fig 4: Plot between J, the total evaporative flux (in m3/s) and the contact angle (radians). The tendency to attain
infinity at certain angles is reflective of the singularity attained as the contact angle may be made to tend to 90
degrees or odd multiples thereof.

Fig 5: The variation of the Time required for evaporation of a given droplet volume with the contact angle
made on the substrate. The almost linear decreasing trend is an affirmation of the earlier point of
evaporation rates being faster for higher initial contact angles.

Now, the temporal evolution of the contact angle is analysed. The expression obtained as Eqn (16) is
plotted to obtain a visualisation, shown in Fig(5). It is seen that there is an almost linearly decreasing

trend to be observed, with the implication that a lower contact angle would necessitate a greater
evaporation time for a given droplet volume. This is in full agreement with the result presented earlier
about the evaporation rate being enhanced at higher contact angles, and stands well-corroborated by
experimental data sets obtained by Hu and Larson [12][13].

Finally, the dynamic evolution of the temperature field on the droplet surface and its relationships with
the initial contact angle and the radial location is studied. The visualisation of eqn (17) presented
above is the source of the analysis. As a matter of plotting the results, the temperature difference from
the initial bulk temperature is chosen to be shown with the initial contact angle in Fig(6). A hyperbolic
decrement is observed, with the maximum decreases being observed for higher contact angles as
expected. The explanation is the higher rate of evaporation at higher contact angles would imply the
requirement of greater heat removal and hence a greater lowering of temperature measured anywhere
in the bulk, although the centre of the surface is chosen here.

Fig(6): A variation of the net decrement of temperature during evaporation with contact angle.
The higher the contact angle initially, the greater the temperature decrease measured at the
centre of the droplet surface.
For the Stokes flow solutions, we obtained the stream function as a function of the radial and angular
directions, and essentially observed the variation of the same. The velocity field as a function of the radial
distance from the centre to the periphery is shown, the velocity at the centre is obviously zero by the
requirements of axisymmetry, but quickly increases to a maximum as the radially outward mass flow rate
is effected by the evaporation, and then again asymptotically decreases to a minimum near the substrate
towards the contact line. The velocity axis is in micrometers per second, and the radius is in millimetres

The trend for the velocity variation (the z-axis, in micrometers/second) in this 3D plot captures the
essential physics both spatially and temporally. While it is understood that the velocity rises from being
zero at the centre to a maximum and falls to a zero at a certain distance as the deposition occurs on the
substrate, the more pertinent observation here is the fact the overall velocity shows a decrement with time,
which is physically explained by the fact that the viscosity shows a net increase with time due to
increasing solutal concentration as the fluid is lost through evaporation, leading to increased viscous drag
which reduces the velocity.
CONCLUSION
The evaporation of a droplet was studied here under the conditions of a concentration limited diffusion
which essentially led to a solution of the Laplace equation in spherical coordinates from which the
evaporative flux and the total flux were derived. The dynamic contact angles variation with respect to
the total flux, the temporal evolution of the contact angle, the temperature variation with respect to the
contact angle and radius are evaluated. The presence of the singularity hints at the unresolved contact
line anomalies about the concentration-limited diffusion approach, with the result of a theoretically
infinite evaporation at edges. The variation of contact angle and temperature is representative of the
dynamic interplay of the surface tension effects in relation to other surface properties as the
temperature conditions are altered, and consequently we obtain the variation as shown herein.
Following the analysis of the evaporation regime itself, we analysed the flow field inside the droplet,
to the physically consistent descriptions of velocity fields inside the droplet, as affected under the
consideration of thermal Marangoni flows. It was established that the velocity rises from a zero at the
centre to a maximum and finally reduces towards the deposition zones near the edges.. Additionally,
the effect of temporal decrease of velocities points to the increased viscous drag as overall
concentration increases as the fluid is lost by evaporation.

REFERENCES
[1] . Marn, H. Gelderblom, D. Lohse, and J. H. Snoeijer. Order-to-Disorder Transition in Ring-Shaped Colloidal
Stains. Phys. Rev. Lett., 107(8):1-4, August 2011.
[2] R. D. Deegan, O. Bakajin, and T. F. Dupont. Capillary flow as the cause of ring stains from dried liquid drops.
Nature, 389(6653):827-829, 1997.
[3] E. R. Dufresne, E. I. Corwin, N. A. Greenblatt, J. Ashmore, D. Y. Wang, A. D. Dinsmore, J. X. Cheng, X. S. Xie, J.
W. Hutchinson, and D. A. Weitz. Flow and Fracture in Drying Nanoparticle Suspensions. Phys. Rev. Lett., 91(22):1-4,
November 2003.
[4] T. P. Bigioni, X. Lin, T. T. Nguyen, E. I. Corwin, T. A. Witten, and H. M. Jaeger. Kinetically driven self assembly of
highly ordered nanoparticle monolayers. Nature Mater., 5(4):265_70, April 2006.
[5Bloemen, O., Flow near the contact line of an evaporating droplet, Doctoral dissertation, January 2012.
[6] P. J. Yunker, T. Still, M. A. Lohr, and A. G. Yodh. Suppression of the coffee-ring effect by shape-dependent
capillary interactions. Nature, 476(7360):308-311, August 2011.
[7] H. B. Eral, D. Mampallil Augustine, M. H. G. Duits, and F. Mugele. Suppressing the coffee stain effect: how to
control colloidal self-assembly in evaporating drops using electrowetting. Soft Matter, 7(10):4954, 2011.
[8] J. Jing, J. Reed, J. Huang, X. Hu, V. Clarke, J. Edington, D. Housman, T. S. Anantharaman, E. J. Huff, B. Mishra, B.
Porter, A. Shenker, E. Wolfson, C. Hiort, R. Kantor, C. Aston, and D. C. Schwartz. Automated high resolution optical
mapping using arrayed, fluid-fixed DNA molecules. Proc. Nat. Acad. Sci., 95(14):8046-51, July 1998.
[9] R. D. Deegan. Pattern formation in drying drops. Phys. Rev. E, 61(1):475-85, January 2000.
[10] Y. Popov. Evaporative deposition patterns: Spatial dimensions of the deposit. Phys. Rev. E, 71(3):117, 2005.
[11] H. Gelderblom, . Marn, H. Nair, A. van Houselt, L. Lefferts, J. H. Snoeijer, and D. Lohse.
How water droplets evaporate on a superhydrophobic substrate. Phys. Rev. E, 83(2):1-6, February 2011.
[12] R. D. Deegan, O. Bakajin, T. F. Dupont, G. Huber, S. R. Nagel, and T. A. Witten. Contact line deposits in an
evaporating drop. Phys. Rev. E, 62(1 Pt B):75665, July 2000.
[13] B. J. Fischer. Particle Convection in an Evaporating Colloidal Droplet. Langmuir, 18(1): 6067, January 2002.
[14] H. Masoud and J. D. Felske. Analytical solution for Stokes flow inside an evaporating sessile drop: Spherical and
cylindrical cap shapes. Phys. Fluids, 21(4):042102, 2009.
[15] A. J. Petsi and V. N. Burganos. Stokes flow inside an evaporating liquid line for any contact angle. Phys. Rev. E,
78(3):1-9, September 2008.
[16] H Hu and R. G. Larson.Evaporation of a sessile droplet on a substrate.J. Phys. Chem. B,106(6):1334-1344, 2002.
[17] H. Hu and R. G. Larson. Analysis of the Microfluid Flow in an Evaporating Sessile Droplet. Langmuir, 21(32):3963-3971, September 2005.
[18] A. Cazabat and G. Guna. Evaporation of macroscopic sessile droplets. Soft Matter, 6(12): 2591, 2010.
[19] J. Eggers and L. M. Pismen. Nonlocal description of evaporating drops. Phys. Fluids, 22(11): 112101, 2010.
[20] J. Polking, A. Boggess, and D. Arnold. Differential Equations with Boundary Value Problems. Prentice Hall, 2nd
edition, 2006. ISBN 9780131862364.

Você também pode gostar