Sunteți pe pagina 1din 9

Applied Thermal Engineering 93 (2016) 214–222

Contents lists available at ScienceDirect

Applied Thermal Engineering


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / a p t h e r m e n g

Research Paper

Water droplets evaporating on horizontal semi-infinite solids


at room temperature
Torgrim Log *
Department of Engineering, Stord/Haugesund University College, Bjoernsonsgt. 45, 5528 Haugesund, Norway

H I G H L I G H T S

• Water droplets drying on semi-infinite solids in stagnant air are studied.


• Water vapour driven convection increases the drying rate for real water droplets.
• Heat radiation dominates the heat transfer for insulating substrates.
• Heat conduction dominates the heat transfer for non-insulating substrates.
• Thermal conductivities above 1 W/m K only marginally increase drying rate.

A R T I C L E I N F O A B S T R A C T

Article history: A global approach to the evaporating droplets on solid substrates in stagnant air is presented. The focus
Received 2 March 2015 is on simplifying the fundamental physics and mathematics involved in order to develop a sound engi-
Accepted 20 September 2015 neering method for determining droplet temperatures and evaporation rates of sessile water droplets
Available online 9 October 2015
on semi-infinite substrates. A mathematically based factor, γ , is introduced to describe the droplet shape,
where γ takes the value of 1 for flat (pan cake shaped) droplets and 2 for half sphere shaped droplets,
Keywords:
and a representative value in the range of 1–2 between these extremes. It is shown that droplets on ther-
Evaporation
mally insulating substrates gets colder and evaporates at about half the rate of similar droplets on high
Drying
Droplet thermal conductivity substrates. Heat radiation dominates the evaporation rate for insulating sub-
Sessile strates while substrate thermal conductivities above 1 W/m K do not significantly increase the evaporation
Semi-infinite solid rate. The theoretical findings were verified by recording temperature and drying time of droplets evapo-
rating on semi-infinite substrates of XPS foam, PMMA and stainless steel at 24 °C and about 30% relative
humidity. The convective water vapour flow from the cold droplets was directed upwards due to the lower
molar mass of the water vapour. For the 10–13 mm radius droplets studied in the present work, the average
convective radius to ambient air was about 15 mm ± 5 mm larger than the droplet radius. It was largest
for droplets on stainless steel and smallest for droplets on the XPS-foam.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction mechanisms involved in droplet evaporation, including internal


surface driven and density driven flows, are complicated. During
Evaporation processes are of general interest for the agricultur- the past years, new knowledge has been developed on the evapo-
al [1], industrial [2] and academic [3] community, as well as for ration of sessile droplets, often involving complex calculations. The
individuals [4]. Several parameters come into play as water drop- droplet evaporation may be divided into three separate modes, i.e.
lets hit objects, wet the surface in different ways, and start constant contact area, constant contact angle and a mixed mode in-
evaporating. Droplet volumes and contact angles play a role [5] and dependent of the initial quantity of water droplets and the
may depend on impurities, additives, etc. In firefighting, additives hydrophobic properties of the substrate [9]. Some of this new knowl-
are used to reduce surface tension for improved performance [6]. edge may, however, become handy when considering simplifications.
The substrate material’s thermal properties [7] as well as the tem- The strange observation of plastic mugs and plates not drying
perature and ambient air relative humidity also come into play, but well in the modern dish washer, while the ceramic mugs and plates
to a less extent for substrates at elevated temperatures [8]. The dry far better, caught the attention of the author as well as others
[4]. And why do the high thermal conductivity stainless steel spoons,
knives and forks not dry significantly faster than the ceramic objects?
* Tel.: +47 900 500 01; fax: +47 5270 2601. The evaporation of volatile liquids on thin substrates has been studied
E-mail address: torgrim.log@hsh.no. by several researchers, both with respect to the drying process [10],

http://dx.doi.org/10.1016/j.applthermaleng.2015.09.108
1359-4311/© 2015 Elsevier Ltd. All rights reserved.
T. Log/Applied Thermal Engineering 93 (2016) 214–222 215

ΔHevap , much heat is required from somewhere for the evapora-


tion process to continue. The heat may to a large extent be supplied
by the solid substrate material. Studying droplet evaporation rates
as a function of substrate thermal properties is therefore of theo-
retical as well as practical and engineering interest.
Especially for droplets on insulating substrates, the droplet may
cool significantly due to the limited heat supply for the entropy
driven evaporation process. It would then generally be assumed that
the stagnant air close to the droplet would cool and become denser
and therefore start moving downwards due to the ideal gas law:

PV = nR g T ( J) (3)

where P (N m−2) is the pressure, V (m3) is the volume, n (mol) is


the number of moles, R g (8.314 J/mol K) is the molar gas constant
and T (K) is the absolute temperature.
The water vapour has, however, due to the low molar mass
(0.01802 kg/mol), significantly lower density than the ambient air
(molar mass 0.02897 kg/mol):
Fig. 1. Principal sketch for the evaporating droplet setup (not to scale). TCK – type PM gas
K thermocouple, diameter 0.5 mm. gas =
Rg T
(kg m3 ) (4)

Compared to air, the molar mass of the water vapour is 37.8%


stain formation [11], nanofluid droplets [12], and even for ice for-
lower. Even though the water vapour is a few K colder than the air,
mation [13]. A good overview on the evaporation of sessile droplets
the water vapour density will still be lower than the ambient air,
is given by Kovalchuk et al. [14] and Erbil [15]. The focus is usually
i.e. the buoyancy forces acting on the water vapour will be di-
on microdroplets on heated substrates [16,17]. No research on evap-
rected upwards. The water vapour pressure is only a few % of the
oration of volatile liquid droplets on thermally thick substrates was,
ambient pressure. The relative density of the layer close to the droplet
however, found in the literature.
surface will not be that much lower than the ambient air, i.e. a larger
The purpose of the present work is, to an engineering accura-
volume is involved for this buoyancy. The vapour released will then
cy, to predict droplet temperature and evaporation rates for sessile
make an upwards directed density driven plume of temperature a
droplets on semi-infinite substrates ranging from insulating ma-
little lower than the ambient temperature.
terials to metals. Any simplifications should be based on sound
The saturated vapour pressure of water may be calculated by a
engineering mathematics and physics, and the calculations easily
three-parameter equation [18]:
handled in a spread sheet. The novelty in the present work is related
to the mathematical simplifications of such semi-infinite systems
⎛ 101325 ⎞
as well as the understanding that may be achieved from pure math- Psat = ⎜
⎝ 760 ⋅ 0.9959 ⎟⎠
⋅ 10(8.07131−1730.63 (−39.724+T )) (N m2 ) (5)
ematical solutions when the droplets evaporate into completely
stagnant air. The heat transfer mechanisms and their role in the heat The factor 0.9959 was introduced to make the vapour pres-
transfer are explained for this ideal case. For the evaporation of real sures fit best possible to the 1990 International Temperature Scale
droplets, a simple convective radius is introduced to explain devi- giving less than 0.03% error for the calculated vapour pressure in
ation from the simplified mathematical diffusion model due to the temperature range of interest in the present work, i.e. 18–
convective water vapour and air currents. The Bernoulli principle 24 °C. The concentration may be calculated from Eq. (4). The ambient
is used to discuss the influence of convection for different evapo- water vapour concentration may be calculated from the ambient
ration rates, i.e. droplets on substrates of different thermal properties. temperature and the recorded relative humidity fraction (RH):

2. Water vapour C 0 = RH ⋅ C (T ) (kg m3 ) (6)

The process involved when water droplets evaporate may be ex- where C (T ) (kg/m3) is the saturation vapour concentration (density)
pressed by: at a given temperature. The concentration difference between vapour
just at the droplet surface at temperature T and the ambient vapour
H 2O(l ) = H 2O( g ) (mol) (1) concentration, i.e. ( C (T ) − C 0 ), represents the driving force for diffu-
sion of water vapour from the droplet surface. The heat flow required
with the associated Gibbs energy change: for maintaining a given evaporation rate may be expressed
by:
ΔG = ΔHevap − T ⋅ ΔSevap ( J mol) (2)
Q evap = m
 ΔHevap (W ) (7)
where ΔHevap (J/mol) is the enthalpy change (heat of evaporation),
T (K) is the temperature and ΔSevap (J/mol K) is the entropy
change. ΔHevap (2.444 MJ/kg) is a significant positive value due to
the chemical properties of water. A large entropy production is there- 3. Heat transfer to the droplet
fore the driving force of the evaporation process to make ΔG < 0 .
It is especially large due to the small molar mass of water The heat flux to an evaporating droplet may consist of all the
( M w = 0.01802 kg mol ) resulting in many vapour molecules per three principal forms of heat transfer, i.e. heat transfer by conduc-
unit mass evaporated. This implies that given a situation where the tion from the solid substrate (Fourier’s law):
whole system is initially at temperature T0 , as shown in Fig. 1, the
droplet will get cooled as it starts evaporating. Due to the large Q k′′ = − k ⋅ dT dx (W m2 ) , (8)
216 T. Log/Applied Thermal Engineering 93 (2016) 214–222

heat transfer by convection: ing into stagnant air at room temperature. Any set-up of internal
convection will even further reduce the internal temperature gradients.
k
Q h′′ = a ⋅ ΔT = h ⋅ ΔT (W m2 ) , (9) After the short initial period, where most heat to the evapora-
δa tion is supplied by the internal droplet heat capacity, the droplet
acts as a disc shaped negative heat source on the surface of the sub-
and heat transfer by radiation: strate. Assuming now that the evaporation rate is constant, and the
substrate is a semi-infinite solid, the solution to the mathematical
Q rad
′′ = ε ⋅ σ ⋅ T 4 (W m2 ) , (10)
problem of heating (or cooling) through a circular surface may be
found in the literature [21]:
where k and ka (W/m K) are the thermal conductivities of the sub-
strate and air, respectively, dT dx (K/m) is the temperature gradient,
ks (T − T0 ) ⎧⎪ ⎡ ⎛ rd ⎞ ⎤ ⎪⎫
2
h (W/m2 K) is the convective heat transfer coefficient, ε (≤1) is = z − ierfc ⎢ Fo*
s ⎨ierfc (Fo*) z 1+ ⎜ ⎥
⎝ z ⎠ ⎥ ⎬⎪
2 Fo ⎟ (14)
the emissivity, σ (5.67·10−8 W/m2 K4) is the Stefan–Boltzmann con- Q ′′rd ⎪⎩ ⎢⎣ ⎦⎭
stant and T (K) is the absolute temperature of the radiating object.
The droplet will radiate back to the surroundings with its own emis- where rd (m) in the present work is the droplet radius and z (m)
sivity. For the radiation wavelengths and droplet thickness of interest is the centerline depth into the substrate layer. The Fourier number
in the present work, the droplet emissivity may be assumed unity of the solid substrate is given by:
[19]. The emissivity of the surroundings and the reflection at the
droplet surface are the parameters influencing the resultant emis- a st
Fos = (15)
sivity of radiation exchange between the droplet and the rd2
surroundings.
Ignoring the heat capacity of the droplet, t (s) is the time from
The droplets may have different shapes depending on their size.
applying the droplet to the substrate. The modified Fourier number
As an example, due to the surface tension, a small droplet may look
is given by:
like a half sphere. A larger droplet takes a flatter “frisbee” shape.
Both these droplet shapes often get a circular footprint. In the present 1
work, only such circular footprint droplets and constant contact area Fo*z = (16)
2 Foz
are considered for the simplicity of the geometrical shape as well
as this, for many substrates, represents the classical shape of a sessile where the Fourier number based on the depth z is given by:
droplet. Heat transfer to this circular shape therefore needs to be
investigated. a st
Foz = (17)
z2
4. Substrate heat transfer to flat water droplets The integral error function is defined by:

Since sessile water droplets at horizontal surfaces in general are ierfc (ξ ) ≡ −ξerfc (ξ ) + exp ( −ξ 2 ) π (18)
quite axis symmetric, the heat equation in cylindrical coordinates
may be used to describe the heat transfer from the substrate ma- and the complementary error function is defined by:
terial to the evaporating droplet:
2 φ
erfc (ϕ ) ≡ 1 − ∫
2
e −η dη (19)
1∂⎛ ∂T ⎞ ∂ ⎛ ∂T ⎞ ∂T π
(W )
0
⎜ ksr ⎟⎠ + ⎜⎝ ks ⎟⎠ = sC Ps
3
m (11)
r ∂r ⎝ ∂r ∂z ∂z ∂z
Generally, a finite negative heat source cannot cool down an in-
where r (m) is the radius, z (m) is the height, T (K) is the tem- finite substrate material infinitely. The temperature of the circular
perature, ks (W/m K) is the substrate thermal conductivity, s (kg/m3) plane negative heat source will then finally approach the limiting
is the substrate density and C Ps (J/kg K) is the substrate specific heat value and the temperature drop of the surface just below the droplet
capacity. Since temperature changes of only a few K are expected will not develop any further. For z → 0 , Eq. (14) is simplified to:
in the substrate material, the thermal properties of the substrate
ks (T − T0 )
will stay constant in the temperature range in the present work. = f pl ( Fos ) (20)
Eq. (11) may then be simplified: Q ′′rd

∂T ⎧ 1 ∂ ⎛ ∂T ⎞ ∂ 2T ⎫ where f pl ( Fos ) is given by:


= as ⎨ ⎜r ⎟ + 2 ⎬ (K s) (12)
∂t ⎩ r ∂r ⎝ ∂r ⎠ ∂z ⎭
⎧ 1 ⎛ 1 ⎞⎫
f pl ( Fos ) = 2 Fos ⎨ − ierfc ⎜ ⎬ (21)
where as (m2/s) is the substrate thermal diffusivity given by: ⎩ π ⎝ 2 Fos ⎟⎠ ⎭

ks The heat flux is given by:


as =
sC Ps
(m2 s) (13)
k (T − T0 )
Q s′′= s f pl ( Fos ) ( W m2 )
−1
(22)
A first assumption is that the temperature gradients within the rd
droplet are small compared to the temperature difference between
average droplet temperature and ambient temperature. This holds The function f pl ( Fos ) is shown in Fig. 2. For large Fos , the func-
well for a droplet evaporating slowly on an insulating horizontal sub- tion f pl ( Fos ) approaches unity and the heat flow to the droplet at
strate [20] where the temperature gradients after the short initial temperature Td is then simply given by:
period will mainly be established in the depth of the insulating sub-
strate material. It will be shown later on that the droplets in the Q s = π rd ks (Td − T0 ) (W ) (23)
present work do not evaporate that much faster on a conducting sub-
strate, i.e. the error due to internal resistance to heat transfer is If a layer is required between the droplet and the solid sub-
negligible regardless of the substrate material for droplet evaporat- strate to make proper wetting angles, this layer of thickness δ i (m)
T. Log/Applied Thermal Engineering 93 (2016) 214–222 217

Daw (C (T ) − C 0 )
 ′′ =
m
rd
f pl ( Fom )
−1
(kg m2 s ) (28)

where the mass diffusion Fourier number, Fom , is given by:

Dawt
Fom = (29)
rd2

The evaporation rate is given by:

 = π rd Daw (C (T ) − C 0 ) f pl ( Fom )−1


m (kg s) (30)

The function f pl ( Fom ) as a function of Fom and the relative


values of f pl ( Fos ) f pl ( Fom ) for representative substrates are shown
in Fig. 2. It is clearly seen that the heat transfer to the droplet from
the aluminum substrate and the stainless steel substrate quickly es-
tablishes quite steady heat flow, i.e. from Fom ≈ 5, which
Fig. 2. The function f pl (Fom ) as a function of Fom and the relative values corresponds to ~100 s. At Fom = 15, which corresponds to 900 s (15
of f pl (Fom ) f pl (Fos ) for representative substrates: a = 5 ⋅ 10−5 m2 s (≈ a Aluminium ) ;
minutes), even the XPS-foam has nearly reached a steady state heat
a = 4 ⋅ 10−6 m2 s (≈ aStainlessSteel ) ; a = 1⋅ 10−6 m2 s (≈ a XPS − foam ); a = 1⋅ 10−7 m2 s (≈ aPMMA ).
flow, while the PMMA substrate still needs more time to reach steady
heat flow.
For large Fom (and Fos ) the mass flux is simply given by:

and thermal conductivity ki (W/m K) will be in series with the sub-  = π rd Daw (C (T ) − C 0 )
m (kg s) (31)
strate. Eq. (23) must then be changed accordingly:
The heat consumed by evaporation is then given by:
−1
⎛r δ ⎞
Q s = π rd2 (T − T0 ) ⎜ d + i ⎟ (W ) (24) Q evap = π rd Daw (C (T ) − C 0 ) ΔHevap (W ) (32)
⎝ ks ki ⎠

The temperature drop over this layer is given by:


6. Mass flux from small (half sphere) droplets
Q sδ i
ΔTi = (K ) (25)
π rd2ki Using the heat equation for spherical coordinates, and apply-
ing the solution from Carslaw and Jaeger [23], the mass flux from
For very short times, just after the droplet is placed on the sub- the half sphere droplet to stagnant air is given by:
strate, i.e. Ford → 0, Eq. (21) simplifies to:
 = 2π rd Daw (C (T ) − C 0 ) f sp ( Fom )−1
m (kg s) (33)
2
f pl ( Fos ) = Fos (26)
π where

ks (
f sp ( Fom ) = erfc ⎡⎣1 2 Fom ⎤⎦ ) (34)
The heat flow is then proportional to , i.e. the thermal
as
The function f sp ( Fom ) is shown in Fig. 3. For large Fom , the mass
effusivity, which may be recognized as ks ρsCps . In the present study,
flux for pure diffusion from the half sphere droplet simplifies to:
early in the droplet evaporation phase, the heat flow is far from con-
stant due to the decreasing droplet temperature, etc. For other  = 2π rd Daw (C (T ) − C 0 )
m (kg s) (35)
geometries, such as thin substrates where the length scale in the
Fos goes radially out from the center of the droplet, the effusivity
may, however, become the dominant heat transfer parameter to the
solid substrate [4].

5. Mass flux from large (flat) droplets

The diffusion of water vapour is described by the Fick’s law:

 ′′ = − Daw ⋅ dC dx
m (kg s m2 ) (27)

where Daw (m2/s) is the intermolecular diffusion coefficient of water


vapour in air and dC dx (kg/m3 m) is the concentration gradient.
At the temperatures of interest in the present work, the intermo-
lecular diffusion coefficient is 25·10−6 m2/s [22]. Fick’s law (Eq. 27)
is in principle the same type of equation as the Fourier law of heat
conduction (Eq. 8). The mathematics involved in the heat transfer
from the solid to the flat pan cake shaped droplet may therefore
be used also for describing the diffusion process. Similar to tran-
sient heat transfer in the substrate in Eq. (22), the mass flux of vapour
into stagnant air is given by: Fig. 3. The function f sp (Fom ) as a function of Fom .
218 T. Log/Applied Thermal Engineering 93 (2016) 214–222

Comparing Eq. (35) to Eq. (31), it is seen that the mass flux from The heat supply needed for the evaporation is then given by:
the half sphere droplet is 2 times the mass flux form the thin plane
⎛1 1⎞
droplet. For cm size water droplets evaporating at heated sur- Q evap = γπ Daw (C (T ) − C 0 ) ΔHevap ⎜ − ⎟ (kg s) (44)
faces, Nakoryakov et al. [24] discussed introducing a dimensional ⎝ rd Ra ⎠
droplet shape factor dependent on evaporation time to the initial
diameter of droplet. In the present work, however, introducing a At constant evaporation rate (i.e. large Fourier numbers), the re-
dimensionless constant γ for the droplet shape, where γ takes lationship between the convective heat flow to the droplet and the
the value of 1 for the plane droplet and 2 for the half spherical heat flow for evaporation is given by:
droplet is justified mathematically. Any droplet shape in between
Q ka (T − T0 ) ⎛1 1⎞
these extremes may be represented by a droplet shape factor γ in ξ=  h = ⎜ − ⎟ (45)
the range of 1–2. Contact angles > 90°, i.e. hydrophobic substrates, Q evap Daw (C (T ) − C 0 ) ΔHevap ⎝ rd Ra ⎠
will not be evaluated. A sound estimate of the droplet shape
The heat flow from the fresh air layer towards the droplet ex-
factor may be based on the height to radius relationship of the
periences a counter flow of cold water vapour from the droplet
droplet, i.e.
reducing the influence of convective heating of the droplet. Nu-
γ = hd rd + 1 (36) merical modelling did, however, show that the counter flow effect
is quite marginal compared to the convective heat flow to the droplet
The mass flux is then given by: at the conditions studied in the present work.

 = γπ rd Daw (C (T ) − C 0 )
m (kg s) (37)
8. Heat balance
Smaller droplets generally show a half sphere shape due
to the surface tension, i.e. γ = 2 . This leads to the conclusion During the first short period of droplet cooling, heat is mainly
that smaller droplets generally achieve lower temperatures supplied by the internal heat capacity of the droplet. When this heat
than larger (flat) droplets on the same semi-infinite substrate supply is exhausted, heat must be supplied from elsewhere by the
material. three principle heat transfer mechanisms. The heat consumed is
given by Eq. (44). The heat flows to the droplet by substrate con-
duction and ambient air conduction/convection are, respectively,
7. Convection given by Eqs. (24) and 25. The heat flow by radiation is given by:

As a first approximation, it is assumed that the air is com- Q rad = γπ rd2εσ (T04 − Td4 ) (W ) (46)
pletely stagnant. Based on the diffusion of mass in Eqs. (31) and (35),
the heat flow from a sessile spherical droplet and flat droplet to the where the emissivity in the present work is estimated to 0.8. The
stagnant air is, respectively, given by: droplet temperature will stabilize at a temperature ensuring balance
in these heat flows, i.e.
Q a ,sp = 2π rd ka (T0 − Td ) f sp ( Foa ) (W )
−1
(38)
( )
Q evap − Q s + Q a + Q rad = 0 (W ) (47)
and
The saturated vapour pressure described by Eq. (5) and the equa-
Q a , pl = π rd ka (T0 − Td ) f pl ( Foa ) (W )
−1
(39) tions involved for describing the involved heat flows and heat
consumption are easily programmed into a spread sheet. A solver
where function can then be used to find the droplet temperature satisfy-
ing Eq. (47).
aat
Foa = (40) It is worthwhile commenting that the heat flow by radiation,
rd2
Q rad , is a function of the droplet radius squared while the other
where aa (m2/s) is the thermal diffusivity of air and ka (W/m K) is factors in Eq. (47) are linearly dependent on the droplet radius, or
the thermal conductivity of air. For the temperatures in the present a combination of rd and Ra . This implies that the radiant contribu-
work, these parameters take the values of 19·10−6 m2/s and 0.026 W/ tion increases more with droplet size than the other heat transfer
m K, respectively [25]. mechanisms and may become the dominant heat transfer param-
For large Foa , the heat flow from the stagnant air to the droplet eter, especially for droplets on insulating substrates. This is in
simplifies to: agreement with the study by Jang et al. [26] regarding evapora-
tion of 100 μm diameter sessile water droplets in an infrared
Q a = γπ rd ka (T0 − Td ) (W ) (41) radiation field. In the present work, it is also seen that conduction
to the substrate does not depend on the droplet shape factor γ .
Assuming that there is a buoyancy driven flow in the vicinity of The most simple case is, however, when the droplet is com-
the droplet and that air of ambient properties are supplied to the pletely flat, i.e. γ = 1 and there is no convection, i.e. Ra → ∞ and
−1
droplets at radius Ra , the steady state heat flow from the ambient ⎛1 1⎞
− → rd . The results for a droplet of radius of 0.01 m under
air to the droplet is given by: ⎝⎜ rd Ra ⎠⎟
such conditions are given in Table 1. It may be noted that since the
⎛1 1⎞
Q a = γπ rd ka (T0 − Td ) ⎜ − ⎟ (W ) (42) whole system is at ambient temperature T0 , this temperature is also
⎝ rd Ra ⎠ the dry bulb temperature.
It may be quite surprising that increasing the thermal conduc-
The diffusive mass flow from the droplet to this ambient layer tivity more than three orders of magnitude gives less than a doubling
is similarly given by: in evaporation rate. There is in fact not much increase in evapora-
tion rate from the Borosilicate glass (Pyrex) ( kws ≈ 1 W m K ) to the
 = γπ Daw (C (T ) − C 0 ) ⎛⎜ 1 − 1 ⎞⎟
m (kg s) (43) metals. A thermal conductivity around this level therefore seems
⎝ rd Ra ⎠ to be the turning point when further improvement in thermal
T. Log/Applied Thermal Engineering 93 (2016) 214–222 219

Table 1
Calculated heat fluxes for a completely flat 10 mm radius droplet on semi-infinite XPS-foam, PMMA, Pyrex glass, stainless steel (AISI316) and aluminum substrates evapo-
rating in stagnant air ( T0 = 24°C and RH = 50%, i.e. a humidity ratio (kg water/kg air) of 0.0092).

ks (W/m K) Q s (W) Q rad (W) Q a (W)  (kg/s)


m T − T0 (K)
−9
XPS-foam 0.03 0.0037 0.0057 0.0032 5.1·10 3.88
PMMA 0.19 0.0119 0.0029 0.0016 6.7·10−9 1.99
Pyrex glass 1.13 0.0185 0.0008 0.0004 8.1·10−9 0.52
AISI316 14.5 0.0207 0.0001 <10−4 8.5·10−9 0.045
Aluminum 140 0.0209 <10−4 <10−4 8.6·10−9 0.005

conductivity barely increases the evaporation rate. The explana- too large at the end of the drying period. A similar simplification
tion for this is that the droplet temperature is kept fairly close to gives an average radius rd = r0d − h0d 2 . Except in the early phase of
the ambient temperature even for substrates of intermediate thermal droplet evaporation, the droplet temperature does not change much
conductivity like the Borosilicate glass. Increasing the thermal con- during the major part of the evaporation period. An average value
ductivity beyond this level cannot do more than marginally increase of the concentrations differences may therefore be used, i.e. ΔC rep-
the droplet temperature, with ambient temperature as the limit. resenting the average of C (T ) − C 0 during the evaporation process.
For a sessile droplet set directly on a high thermal conductivity The data satisfying the heat balance may then be used to cal-
material like stainless steel, i.e. ks > 10 W m K , where the droplet culate the droplet evaporation time. Checking this calculation with
temperature remains close to the ambient temperature during the the recorded evaporation time for different radius out to the fresh
evaporation process, Eq. (43) may be simplified to: air layer, i.e. Ra , by trial and error this distance may be estimated.

 = γπ DawC (T0 ) (1 − RH ) ⎛⎜ 1 − 1 ⎞⎟
m (kg s) (48) 10. Experiments
⎝ rd Ra ⎠
Droplets of 0.50 ml (~10 mm radius) were produced with a mi-
Assuming stagnant water within the droplet and no contact re-
cropipette (Socorex Swiss 200–1000 μl) and put on extruded
sistance to the solid substrate, the internal temperature drop may
polystyrene foam (XPS), polymethylmethacrylate (PMMA) and
be estimated by:
AISI316 stainless steel substrates. The thermal conductivity of these
Q shd materials had been measured previously [27]. A sheet of Kapton
ΔTd = (K ) (49) (thickness 25 μm, thermal conductivity 0.12 W/m K) was put
π rd2kw
between the substrate material and the droplet to make the drop-
where hd (m) is the droplet height and kw (W/m K) is the thermal lets wet similarly for each substrate material. According to Eq. (25),
conductivity of water. Assuming a droplet height of 1 mm, radius this thin layer exerts only a minor influence on the thermal trans-
of 0.01 m, kw = 0.60 W m K and typical values of Q s from Table 1 port from the droplet to the substrate. A cage of very fine mesh steel
gives ΔTd = 0.02 K and 0.11 K for XPS-foam and aluminum, respec- screen was set up for making an environment of stagnant air within
tively. The real ΔTd is, however, expected to be smaller due to internal the laboratory. The ambient lab temperature was set to a constant
convection within the droplet due to surface tension driven flow value, the light was turned off and the window shielded to mini-
[7] and density differences. mize any stray heat radiation. The recordings were made only on
The speed of the vapour leaving the pancake shaped flat droplet days with grey weather and minimal solar radiation to the lab build-
is given by the mass flow divided by the area of the droplet times ing. The recorded ambient temperature was therefore the dry bulb
the vapour density leaving the surface. For a droplet evaporating temperature in the droplet drying experiments.
into stagnant air, this simplifies to: Thermocouples of type K (diameter 0.5 mm) were used for the
temperature recordings. The experimental setup is shown in Fig. 1
Daw ⎛ C (T ) − C 0 ⎞ with indication of thermocouple location within the droplet. Since
v= ⎜ ⎟ (m s ) (50)
the droplets would dry some, and get a little too cold when set, the
rd ⎝ C (T ) ⎠
water was heated a few degrees above room temperature before
Given the droplet size, temperatures and ambient relative hu- setting the droplets. A hand held precision psychrometer (type
midity in the present work, the vapour speed is on the order of RH390, Extech Instruments) was used to measure relative humid-
1 mm/s at the droplet surface. ity (RH) prior to, and during, the experiments. Recorded droplet
temperatures for the substrates studied are shown in Fig. 5 and
9. Droplet drying time Table 2.
The temperature above droplets evaporating on PMMA was also
The droplet evaporation may be divided into three separate measured with the 0.5 mm diameter thermocouple. It was dem-
modes, i.e. constant contact area, constant contact angle and a mixed onstrated that the temperature decreased with height over the center
mode independent of the initial quantity of water droplets and the
hydrophobic properties of the substrate [9]. In order to get a handy
engineering evaporation model for sessile droplets, some simpli-
fications have to be made. Assuming now that the droplet has an
initial radius r0d = 10 mm and height h0d = 2 mm , a sound estimate
may be to assume that the droplet has a shape as indicated in Fig. 4.
The simplified air droplet contact area to the footprint area is
then 1.15 ( γ 0 = 1.15 ). Assuming now that the droplet dries such that
both the height and radius changes similarly with time, i.e.
dhd dt = drd dt , the factor γ may be calculated for each time step.
Even easier is to set an average constant factor γ = (1 + γ 0 ) 2 = 1.075 Fig. 4. Simplified droplet shape for evaporation calculations. The solid line repre-
during the drying period, i.e. a little too low in the start and a little sents the simplification. rd , droplet radius; hd , droplet height.
220 T. Log/Applied Thermal Engineering 93 (2016) 214–222

temperature and the more heat is supplied by thermal radiation.


Radiation may simply take over as the most important heat trans-
fer parameter for droplets on insulating substrates in a stagnant air
environment at room temperature. For higher thermal conductiv-
ity substrates, the mass diffusion process is the only limiting factor.
Substrates with thermal conductivity of 1 W/m K is able to keep the
temperature of the evaporating droplet fairly close to ambient tem-
perature. Substrate thermal conductivity above 1 W/m K results in
droplet temperatures approaching the ambient temperature giving
minimal gain in evaporation rates, especially for thermal conduc-
tivities above 10 W/m K. Another factor also contributing to the self-
regulating process is the convective distance from the droplet to fresh
air. In the present work, this distance increases with increasing
thermal conductivity further slowing the mass diffusion from drop-
lets on highly conducting substrates.
Fig. 5. Typical temperature development for the first 15 minutes for 10 mm radius
droplets on the semi-infinite substrates studied at T0 = 24.5°C ± 0.5°C and In buoyancy driven convective heat loss from a plane heat source,
RH = 0.31 ± 0.02, i.e. a humidity ratio of 0.0059 ± 0.05. the heat transfer coefficient generally increases slightly with in-
creasing source strength. This is due to increased buoyancy which
gives increased air speed making the aerodynamic boundary layer
of the droplet. At elevations of 2, 5 and 15 mm along the center line to ambient air, δ a (Eq. 9), smaller by a fraction of the relative in-
above the droplet surface, a temperature reduction of 1.7, 1.0 and crease in upward buoyancy. In the present work, it seems to be the
0.2 °C was respectively recorded. At 2 mm height above the PMMA other way around. The stronger diffusive source (droplet on the stain-
substrate and 5 mm away from the droplet, only 0.2 °C tempera- less steel) gives a longer average distance δ a = Ra − rd to ambient air
ture drop from ambient was recorded. Though the thermocouple than the weaker diffusive sources (droplets on PMMA and XPS-
was too thick to measure temperatures in air properly, the results foam). This puzzling fact may to some extent be explained by the
indicate that the water vapour released at the droplet surface moved influence of increased buoyancy versus increased mass flow. With
upwards due to its lower molar mass. The convective air current increased mass flow, the driving ΔP may increase linearly. This may
around the droplet is therefore an upwards driven buoyant flow re- give an increase in speed proportional to ΔP . If this is a factor in
gardless of the droplet temperature being a few degrees below the the explanation, a doubling in mass flow then gives a 2 increase
ambient temperature. in speed. It may then be the case that the increased mass flow
From the temperature recordings around the droplet, it was es- pushing the distance δ a larger dominates the increased speed in-
tablished that the ambient air close to the droplet was influenced fluence trying to decrease the distance δ a by a factor 1 2 . More
by the buoyancy of the produced vapour. A value of Ra smaller than accurate measurements are needed to verify whether this can be
infinity was therefore aimed at in each calculation until the best a sound explanation. Sobac and Brutin [17] also addressed the need
match in evaporation time for each droplet was achieved. The results for more accurate models to account for the buoyant convection in
are shown in Table 3. the vapour transport.
Some instability was observed in the droplet temperature es- The substrate material close to the droplet is cooled down, es-
pecially for the droplets on stainless steel. This is seen as bumps pecially for the low conductivity substrates. A cold substrate surface
in the temperature trends in Fig. 5. There may also have been similar in contact with the air just outside the droplet surface will cool the
instabilities for droplets on the PMMA and XPS-foam substrates, but air in the vicinity of the droplet. A cold air current along the sub-
lower internal temperature gradients as well as a thick thermo- strate surface directed away from the droplet can therefore not be
couple compared to the droplet height did not allow for any detailed excluded. Since convection represents a significant fraction of the
analysis of this phenomenon. heat transfer to the droplet on the XPS-foam, this mechanism may
also be a part of the explanation of shorter distance to ambient air
11. Discussion for the XPS-foam substrate compared to the more conducting
substrates.
Droplet evaporation turns out to be a self-regulating process. The The measured temperatures for the droplets on XPS-foam, PMMA
lower the substrate thermal conductivity, the lower droplet and stainless steel match the calculated temperatures within 0.2 °C.
The largest relative error was measured for the droplet on the stain-
less steel substrate. The instability in this droplet temperature may
Table 2 indicate that there are some unstable temperature gradients within
Measured temperatures and evaporation time for representative sessile droplets on
the droplet or unstable liquid transport. The slight mismatch in mea-
XPS-foam, PMMA and AISI316 stainless steel.
sured and calculated temperature could therefore have been larger
rd (mm) hd (mm) γ () T0 (°C) RH ( ) ΔTd,∞ (K) t evap (h) or smaller dependent on the thermocouple location. The fact that
XPS-foam 12.5 1.11 1.09 23.3 0.39 5.9 9.4 the droplet surface may have a little lower temperature than the
PMMA 10.2 1.80 1.18 23.3 0.39 3.4 10.1 measured value does not significantly change the conclusions in the
AISI316 11.1 1.45 1.13 22.9 0.38 0.2 7.1
study. It would, however, be worthwhile studying these effects using

Table 3
Best fit Ra matching measured evaporation time for the droplets in Table 2.

Q s (10−3 W) Q rad (10−3 W) Q conv (10−3 W)  (kg/s)


m ΔTd,∞ (K) Ra (mm) Ra − rd (mm)
−8
XPS-foam 7.1 14.8 14.0 1.47·10 6.02 23.9 11.4
PMMA 21.3 6.3 5.5 1.36·10−8 3.50 26.9 16.7
AISI316 46.8 0.2 0.1 1.93·10−8 0.09 31.1 20.0
T. Log/Applied Thermal Engineering 93 (2016) 214–222 221

thinner thermocouples. Then, any temperature gradients within the H Enthalpy (J/mol)
droplet could have been measured. Any rapid temperature changes k Thermal conductivity (W/m K)
within the droplet due to unstable internal flow could maybe also m  Mass flow (kg/s)
have been detected, if any. It is not expected that this instability is m  ′′ Mass flux (kg/m2 s)
due to external instabilities, as observed in adsorption absorption M Molar mass (kg/mol)
situations involving multicomponent atmospheres [28]. Internal in- P Pressure (N/m2)
stabilities have, however, also been found in evaporating FC-72 sessile Q Heat flow (W)
droplets [29] as well as in water droplets evaporating on glass sub- r Radius (m)
strates [30]. The latter work reports instabilities up to 0.2 °C. R Radius (m)
Though the dish washer situation is different with numerous Rg Molar gas constant (J/mol K)
sessile droplets on 0.5–5 mm thick substrates (i.e. knifes to plates) RH Relative humidity (–)
in a high relative humidity forced convection environment, as seen S Entropy (J/mol K)
from the individual droplets, the results in the present work may t Time (s)
to some extent explain why plastic products (k ≈ 0.15 W/m K) dry T Temperature (K)
slowly in the dish washer. It may also indicate why the stainless x Distance (m)
steel forks, knifes and spoons (k ≈ 15 W/m K) do not dry signifi- z Distance (m)
cantly faster than glass and stoneware products (k ≈ 1.5 W/m K) as
they both seem to be above the threshold value where increased Greek symbols
thermal conductivity ceases to contribute in increased evapora- ϕ , η , ξ Dimensionless parameters (–)
tion rates. This is in line with the findings of Sobac and Brutin [17], γ Droplet shape factor (–)
as they, even at substrate temperatures at 75 °C, were not able to δ Thickness (m)
distinguish evaporation rates for 3–4 μl water droplets on heated Δ Change
bronze and copper substrates, i.e. a thermal conductivity range of  Density (kg/m3)
50–400 W/m K. σ Stefan–Boltzmann constant (W/m2 K4)
For thin substrates, more volatile liquids or water at higher sub- ε Emissivity (–)
strate temperatures, where the droplets evaporate significantly faster
than in the present work, the steady state heat transfer may not be Subscripts
achieved. Such transient situations are recently outlined by Maatar 0 Initial condition
et al. [10]. In the present work the true steady heat and mass flows a Air/ambient
were, however, achieved resulting in very simple mathematical d Droplet
solutions. g Gas
i Insulation
12. Conclusion l Liquid
pl Plane
In contradiction to the general belief assuming that air cooled s Substrate
by the water droplet is moving downwards, it is shown that the sp (Half)spherical
transport of water vapour and entrained air is mainly directed sat Saturation
upwards due to the low molar mass of water vapour. Mathemati- w Water
cal solutions based on finite plane sources and semi-infinite solids
give reasonable estimates of droplet temperatures and evapora- Superscripts
tion rates. Based on this knowledge, principle evaporation * Modified
dependencies on substrate material and droplet sizes have been es- ̄ Average
tablished. This is verified for droplets of about 10 mm radius
evaporating on XPS-foam, PMMA and stainless steel. It is shown that
natural convection takes place in the experiments ensuring evap- References
oration rates significantly higher than pure diffusion into completely
stagnant air. [1] G.R. Chegini, B. Ghobadian, Spray dryer parameters for fruit juice drying, World
J. Agric. Sci. 3 (2) (2007) 230–236.
[2] V. Dugas, J. Broutin, E. Souteyrand, Droplet evaporation study applied to
Acknowledgements DNA chip manufacturing, Langmuir 21 (2005) 9130–9136, doi:10.1021/
la050764y.
The work has been partly funded by the Stord/Haugesund Uni- [3] R.D. Deegan, O. Bakajin, T.F. Dupont, G. Huber, S.R. Nagel, T.A. Witten, Capillary
flow as the cause of ring stains from dried liquid drops, Nature 389 (1997)
versity College. The author wants to thank Siles E. Gustafsson for 827–829, doi:10.1038/39827.
supplying the 25 μm thick Kapton sheets used to ensure similar [4] M.-C. Lopes, E. Bonaccurso, T. Gambaryan-Roisman, P. Stephan, Influence of the
droplet wetting for droplets evaporating on different substrate substrate thermal properties on sessile droplet evaporation: effect of transient
heat transport, Colloids Surf. A Physicochem. Eng. Asp. 432 (2013) 64–70
materials. <http://dx.doi.org/10.1016/j.colsurfa.2013.04.017>.
[5] T.A.H. Nguyen, A.V. Nguyen, M.A. Hampton, Z.P. Xu, L. Huang, V. Rudolph,
Nomenclature Theoretical and experimental analysis of droplet evaporation on solid surfaces,
Chem. Eng. Sci. 69 (2013) 522–529, doi:10.1016/j.ces.2011.11.009.
[6] X. Zhou, B. Zhou, X. Jin, Study of fire-extinguishing performance of portable
a Thermal diffusivity (m2/s) water-mist fire extinguisher in historical buildings, J. Cult. Herit. 11 (2010)
C Concentration (kg/m3) 392–397, doi:10.1016/j.culher.2010.03.003.
[7] F. Girard, M. Antoni, S. Faure, A. Steinchen, Numerical study of the evaporating
CP Heat capacity at constant pressure (J/kg K)
dynamics of a sessile water droplet, Micrograv. Sci. Technol. XVIII-3/4 (2006)
D Intermolecular diffusion coefficient (m2/s) 42–46, doi:10.1007/BF02870377.
Fo Fourier number (–) [8] F. Girard, M. Antoni, S. Faure, A. Steinchen, Influence of heating temperature
G Gibbs energy (J/mol) and relative humidity in the evaporation of pinned droplets, Colloids Surf. A
Physicochem. Eng. Asp. 323 (2008) 36–49, doi:10.1016/j.colsurfa.2007.12.022.
h Height (m) [9] J.-H. Kim, S.I. Ahn, H.H. Kim, Evaporation of water droplets on polymer surfaces,
h Heat transfer coefficient (W/m2 K) Langmuir 23 (11) (2007) 6163–6169, doi:10.1021/la0636309.
222 T. Log/Applied Thermal Engineering 93 (2016) 214–222

[10] A. Maatar, S. Chikh, M.A. Saada, L. Tadrist, Transient effects on sessile droplet [20] O.E. Ruiz, W.Z. Black, Evaporation of water droplets placed on a heated
evaporation of volatile liquids, Int. J. Heat Mass Transf. 88 (2015) 212–220, horizontal surface, J. Heat Transf. 124 (5) (2002) 854–863, doi:10.1115/
doi:10.1016/j.ijheatmasstransfer.2015.02.077. 1.1494092.
[11] N. Shahidzadeh, M.F.L. Schut, J. Desarnaud, M. Prat, D. Bonn, Salt stains [21] J.H. VanSant, Conduction Heat Transfer Solutions, UCRL-52863, Lawrence
from evaporating droplets, Nat. Sci. Rep. 5 (2015) 10335, doi:10.1038/ Livermore National Laboratory, 1980.
srep10335. [22] T.R. Marrero, E.A. Mason, Gaseous diffusion coefficients, J. Phys. Chem. Ref. Data
[12] Y.C. Kim, Evaporation of nanofluid droplet on heated surface, Adv. Mech. Eng. 1 (1) (1972) 3–118.
7 (4) (2015) 1–8, doi:10.1177/1687814015578358. [23] H.L. Carslaw, J.C. Jaeger, Conduction of Heat in Solids, second ed., Oxford Science
[13] Y. Wu, X. Zhang, X. Zhang, M. Munyalo, Modeling and experimental Publications, Oxford, 1959. ISBN: 0-19-853368-3.
study of vapor phase-diffusion driven sessile drop evaporation, Appl. [24] V.E. Nakoryakov, S.Y. Misyura, S.L. Elistratov, The behavior of water droplets
Therm. Eng. 70 (1) (2014) 560–564, doi:10.1016/j.applthermaleng.2014 on the heated surface, Int. J. Heat Mass Transf. 55 (2012) 6609–6617,
.05.049. doi:10.1016/j.ijheatmasstransfer.2012.06.069.
[14] N.M. Kovalchuk, A. Trybala, V.M. Starov, Evaporation of sessile droplets, Curr. [25] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, second ed., John
Opin. Colloid Interface Sci. 19 (4) (2014) 336–342, doi:10.1016/ Wiley & Sons, 2001. ISBN 0-471-41077-2.
j.cocis.2014.07.005. [26] J. Jang, C. Lee, J.W. Hahn, Theoretical study on evaporation of sessile water
[15] H.Y. Erbil, Evaporation of pure liquid sessile and spherical suspended drops: droplets on a glass panel with infrared radiation, J. Mech. Sci. Technol. 28 (4)
a review, Adv. Colloid Interface Sci. 170 (2012) 67–86, doi:10.1016/ (2014) 1575–1580, doi:10.1007/s12206-013-1187-3.
j.cis.2011.12.006. [27] T. Log, S.E. Gustafsson, Transient Plane Source (TPS) technique for measuring
[16] S. Semenov, V.M. Starov, R.G. Rubio, Evaporation of pinned sessile microdroplets thermal transport properties of building materials, Fire Mater. 19 (1995) 43–49,
of water on a highly heat-conductive substrate: computer simulations, Eur. doi:10.1002/fam.810190107.
Phys. J. Spec. Top. 219 (31) (2013) 143–154, doi:10.1140/epjst/e2013 [28] A.H. Persad, K. Sefiane, C.A. Ward, Source of temperature and pressure pulsations
-01789-y. during sessile droplet evaporation into multicomponent atmospheres, Langmuir
[17] B. Sobac, D. Brutin, Thermal effects of the substrate on water droplet 29 (43) (2013) 13239–13250, doi:10.1021/la403177r.
evaporation, Phys. Rev. E 86 (2 1) (2012) 021602, doi:10.1103/PhysRevE.86 [29] K. Sefiane, Y. Fukatani, Y. Takata, J. Kim, Thermal patterns and hydrothermal
.021602. waves (HTWs) in volatile drops, Langmuir 29 (31) (2013) 9750–9760,
[18] C. Antoine, Tensions des vapeurs; nouvelle relation entre les tensions et les doi:10.1021/la402247n.
températures, C. R. Séances Acad. Sci. 107 (1888) 836–837, 681–684, 778–780. [30] T.A. Yakhno, O.A. Sanina, M. Volovik, A. Sanin, V.G. Yakhno, Thermographic
[19] T. Log, Radiant heat attenuation in fine water sprays, in: Proc. 7th Int. Fire investigation of the temperature field dynamics at the liquid–air interface in
Science & Engineering Conference, Interflam-96, Cambridge, UK, 1996, pp. drops of water solutions drying on a glass substrate, Tech. Phys. 57 (7) (2012)
425–434. 915–922, doi:10.1134/S1063784212070262.

S-ar putea să vă placă și