You are on page 1of 9

BGHiggins/UCDavis/June_2014/Draft2

Spreading of an Axisymmetric Drop


Background
Consider a axisymmetric drop spreading on a horizontal surface located at z=0. Let the interface of the
drop be axisymmetric such that the interface of the drop can be given by
z = h (r, t) (1)
The flow geometry is shown in the figure below:

z g
h(r, t)

2R(t)

A related problem is the spreading of a spherical drop due to capillary effects only. This is the problem
studied by Blake et al (1997). Hu and Larson have also studied the flow in a spherical drop but in this
case the flow is induced by the evaporation of liquid inside the drop. In their study the contact line was
taken to be pinned.
In the problem considered here we study spreading driven by wetting effects, capillary pressure and
gravity. We examine the shape of the drop after it has reach a condition where its thickness varies
slowly with position. Then the flow is nearly one-dimensional and inertial effects can be neglected. The
flow state is governed by the lubrication approximation. This is the case studied by Lopez et al (1976).
In this report, will show that the evolution equation for h(r,t) that determines the shape of an axisymmet-
ric drop during spreading as a result of capillary and gravity driving forces( see Resnik and Yarin
(2002)) is
∂h 1 ∂ ∂h ∂ 1 ⅆ ⅆh
= rh3 ρ g -− σ r  
∂t 3 μ r ∂r ∂r ∂r r dr ⅆr
In our study we will ignore capillary pressure effects.
The position of the contact line during spreading is given by r=R(t). Note that in the absence of capillary(
wetting) forces the slope of the interface at the contact line is not specified.
For the axisymmetric spreading drop the mean curvature ℋ of the interface is ( see Appendix)
2 SpreadingAxisymetricDrop.nb

ⅆ2 h  dr2 ⅆh /∕ ⅆr


2ℋ= + (2)
3/∕2 1/∕2
1 + (ⅆh /∕ ⅆr)2  r 1 + (ⅆh /∕ ⅆr)2 
In the later stages of spreading we may assume that ⅆh/ⅆr is small so that we can approximate the
curvature as
ⅆ2 h 1 ⅆh 1 ⅆ ⅆh
2H≈ + = r (3)
2 r ⅆr r dr ⅆr
ⅆr

Flow Equations
We will consider the flow problem where the fluid velocity is a function of both r and z. Let the radial
velocity be u(r,z,t) and the axial velocity v(r,z,t). The relevant flow equations ignoring inertial effects are
1 ∂(r u) ∂v
Continuity : + =0
r ∂r ∂z

∂ 1 ∂ ∂2 u ∂P
R -− component of momentum : μ (r u) + = (4)
∂r r ∂r ∂z2 ∂r

1 ∂ ∂v ∂2 v ∂P
z -− component of momentum : μ r + = +ρ g
r ∂r ∂r ∂z2 ∂z
In the later stages of spreading the flow in the drop can be approximating as a lubricating flow such that
the radial component of velocity u(r,t) is much larger in magnitude than the axial component of velocity
v(r,t), i.e.
u (r, t) >> v (r, t)
A formal order of magnitude analysis shows that the flow equations can be approximated as
1 ∂(r u) ∂v
Continuity : + =0
r ∂r ∂z

∂2 u ∂P
R -− component of momentum : μ = (5)
2 ∂r
∂z

∂P
z -− component of momentum : 0= +ρ g
∂z
Integrating the z-component of momentum gives
P (r, z, t) = -−ρ g z + C1 (r, t) (6)

Interfacial Boundary Conditions


For the axisymmetric spreading drop the mean curvature ℋ of the interface is (see Appendix) is given by
ⅆ2 h  dr2 ⅆh /∕ ⅆr
2ℋ= + (7)
2 3/∕2 1/∕2
1 + (ⅆh /∕ ⅆr)  r 1 + (ⅆh /∕ ⅆr)2 
In the later stages of spreading, when the lubrication approximation holds, we may assume that ⅆh/ⅆr is
small so that we can approximate the curvature as
SpreadingAxisymetricDrop.nb 3

In the later stages of spreading, when the lubrication approximation holds, we may assume that ⅆh/ⅆr is
small so that we can approximate the curvature as
ⅆ2 h 1 ⅆh 1 ⅆ ⅆh
2H≈ + = r (8)
ⅆr2 r ⅆr r dr ⅆr
Now at the interface of the drop, the normal stress boundary condition subject to the lubrication approxi-
mation yields the following expression for the pressure field at the interface:
1 ⅆ ⅆh
P (h, t) = PA -− σ r (9)
r dr ⅆr
where PA is the ambient pressure outside the drop. The kinematic condition subject to the lubrication
approximation gives
∂h
= v, at z = h (r, t) (10)
∂t
At the interface z=h(r,t), the vanishing shear stress condition gives (where τ𝜏rr , τ𝜏zr and τ𝜏zz are the stress
components):
hz τrr + τzr 1 -− h2z  -− hz τzz = 0 (11)

Linearizing the above for small slopes ∂𝜕h/∂𝜕z gives


τzr = 0

∂v ∂u (12)
+ = 0, at z = h (r, t)
∂r ∂z
and applying the lubrication approximation, the vanishing of the shear stress at the interface is approxi-
mated by
∂u
= 0, at z = h (r, t) (13)
∂z

Velocity Field in Drop


Using Eq. (5) we can determine the integration function C1 (r, t) in Eq. (6):
1 ⅆ ⅆh
C1 (r,t)= PA + ρ g h -− σ r (14)
r dr ⅆr
Thus the pressure in the spreading drop is given by
1 ⅆ ⅆh
P (r, z, t) = ρ g (h -− z) + PA -− σ r (15)
r dr ⅆr
Note that if we differentiate the pressure field with respect to r we get
∂P ∂h ∂ 1 ⅆ ⅆh
=ρg -− σ r  (16)
∂r ∂r ∂r r dr ⅆr
Thus the radial pressure gradient is independent of z! Thus we can integrate the r-component of veloc-
ity (see Eq. (5) ) , with respect to z, we get
1 ∂P z2
u= + C1 Z + C2 (17)
μ ∂r 2
4 SpreadingAxisymetricDrop.nb

Lubrication Solution
The governing equations we need to solve are
1 ∂(r u) ∂v
+ =0
r ∂r ∂z

1 ∂P z2
u= + C1 Z + C2
μ ∂r 2

∂P ∂h ∂ 1 ⅆ ⅆh (18)
=ρg -− σ r 
∂r ∂r ∂r r dr ⅆr

∂u
= 0, at z = h (r, t)
∂z

u (0, t) = 0, v (0, t) = 0
We begin by integrating the continuity equation to determine an expression for the normal velocity at the
interface
h 1 ∂(r u)
v (h) = -−  ⅆz
o r ∂r
(19)
1 ∂ h
= -− r  u ⅆz
r ∂r o

The radial velocity u(z,t) is given by


1 ∂P z2
u (z, t) = + C1 Z + C2 (20)
μ ∂r 2
Satisfying the no-slip condition z=0 requires that C2 = 0. The zero shear stress condition is satisfied if
∂u 1 ∂P
=0= (h + C1 ) ⟹ C1 = -−h (21)
∂z μ ∂r
Hence we have
1 ∂P z2
u (z, t) = -− h Z (22)
μ ∂r 2
It follows then
h h 1 ∂P z2 1 ∂P h3
 u ⅆz =  -− h Z ⅆz = -− (23)
o o μ ∂r 2 μ ∂r 3
Thus from Eq. (19) and using Eq. (22) we can determine the velocity component v(h):
1 ∂ h 1 ∂ ∂P
v (h) = -− r  u ⅆz = r h3  (24)
r ∂r o 3 μ r ∂r ∂r
Recall that
SpreadingAxisymetricDrop.nb 5

∂P ∂h ∂ 1 ⅆ ⅆh
=ρg -− σ r  (25)
∂r ∂r ∂r r dr ⅆr
Thus
1 ∂ ∂h ∂ 1 ⅆ ⅆh
v (h) = rh3 ρ g -− σ r   (26)
3 μ r ∂r ∂r ∂r r dr ⅆr

Evolution Equation for Spreading Drop


Finally the kinematic condition (se Eq. (10), with Eq. (26)) gives the evolution equation for the free
surface when gravity and capillary forces are important:
∂h 1 ∂ ∂h ∂ 1 ⅆ ⅆh
= rh3 ρ g -− σ r   (27)
∂t 3 μ r ∂r ∂r ∂r r dr ⅆr
If capillary effects are not important then the evolution equation for the free surface is
∂h ρg ∂ ∂h
= r h3 (28)
∂t 3 μ r ∂r ∂r
If gravitational effects are not important, the evolution equation for the free surface is
∂h σ ∂ ∂ 1 ⅆ ⅆh
= -− r h3  r  (29)
∂t 3 μ r ∂r ∂r r dr ⅆr

Evolution Equation when Gravity Forces Dominate


When the spreading of the spherical drop is dominated by gravity forces, the shape of the drop is given
by
∂h ρg ∂ ∂h
= r h3 (30)
∂t 3 μ r ∂r ∂r
subject to the following boundary conditions:
∂h
At r = 0 : = 0,
∂r
(31)
At r -− R (t) : h=0

At t = 0 : h (r, t) = H (R)
We will seek a similarity solution to this problem

Similarity Analysis
It will be convenient for the analysis that follows to introduce a dimensionless time τ𝜏 given by
τ = ρ g t /∕ (3 μ) (32)
Thus the evolution equation becomes
∂h 1 ∂ ∂h
= r h3 (33)
∂τ r ∂r ∂r
6 SpreadingAxisymetricDrop.nb

Next we will look for a similarity solution of the form


r
h (τ, ξ) = τα f (ξ), where ξ = (34)
τβ
In terms of the similarity variables the derivative of h(τ𝜏,ξ𝜉) with respect to τ𝜏 gives
∂h ⅆf ∂ξ
= α τα-−1 f (ξ) + τα
∂τ ⅆξ ∂τ
(35)
ⅆf
= α τα-−1 f (ξ) -− β r τα-−β+1
ⅆξ
The derivative of h(τ𝜏,ξ𝜉) with respect to r gives
∂h ⅆf ∂ξ 1 ⅆf
= τα = τα
∂r ⅆξ ∂r τβ ⅆξ
(36)
ⅆf
= τα-−β
ⅆξ
Substituting Eq. (35) and (36) into the PDE Eq. (33) gives
ⅆf 1 1 ⅆ ∂h
α τα-−1 f (ξ) -− β r τα-−β+1 = ξ h3
ⅆξ 2β ξ ⅆξ ∂ξ
τ

1 1 ⅆ ⅆf
= ξ τ4 α f3 (37)
τ2 β ξ ⅆξ ⅆξ

1 ⅆ ⅆf
-− τ4 α-−2 β ξ f3
ξ ⅆξ ⅆξ
Regrouping terms we get
ⅆf 1 ⅆ ⅆf
τα-−1 α f (ξ) -− β ξ = τ4 α-−2 β ξ f3 (38)
ⅆξ ξ ⅆξ ⅆξ
Thus for f(ξ𝜉) to be independent of τ𝜏- the similarity solution postulate, we must have
α -− 1 = 4 α -− 2 β ⟶ 3 α -− 2 β + 1 = 0 (39)
Consider next the conservation of mass of he drop during spreading ( recall evaporation of liquid in the
drop is assumed to be negligible). Thus the volume of the drop V at any time t is
R (t)
V=2π h (r, t) r ⅆr (40)
0

In terms of the similarity variables this constraint becomes


ξ=Rτβ
α+2 β (41)
V=2πτ  f (ξ) ξ ⅆξ
0

Thus for V to be constant for all τ𝜏, we require that


α + 2 β = 0, and ξ = R (t)  τβ = ξ0 (42)

Solving Eqs. (39) and (42) gives


SpreadingAxisymetricDrop.nb 7

1 1
β= , and α = -− , and ξ0 = R (t)  τ1/∕8 ⟹ R (t) = ξ0 τ1/∕8 (43)
8 4
The similarity equation that must be solve for f(ξ𝜉) such that Eq. (43) is satisfied is
ⅆf 1 ⅆ ⅆf
α f (ξ) -− β ξ = ξ f3 (44)
ⅆξ ξ ⅆξ ⅆξ
By regrouping terms and using Eq. (43) this equation can be expressed as
ξ ξ2 ⅆf ⅆ ⅆf
-− f -− -− ξ f3 =0 (45)
4 8 ⅆξ ⅆξ ⅆξ
A further reorganization gives
1 ⅆ ⅆ ⅆf
ξ2 f -− ξ f3 =0 (46)
8 ⅆξ ⅆξ ⅆξ
Integrating gives
1 ⅆf
ξ2 f -− f3 ξ = K1 (47)
8 ⅆξ
Now at the contact line ξ𝜉 = ξ𝜉0 , but f (ξ𝜉0 ) =0 so that K1 =0. Thus the equation that f(ξ𝜉) must satisfy is
ξ ⅆf
+ f2 = 0, where ξ → ξ0 = Lim r  τ1/∕4 (48)
8 ⅆξ r→R (t)

Integrating gives
3 1/∕3
1/∕8
f (ξ) = ξ20 -− ξ2  (49)
16
where ξ𝜉0 is the position where f(ξ𝜉)=0, i.e., where h(r,t)=0 , i.e., the contact line R(t).
The volume V of the drop is then ( see Eq. (41)) becomes
3 1/∕3 ξ0
1/∕3
V=2π  ξ20 -− ξ2  ξ ⅆξ
16 0

(50)
3 3 1/∕3
4/∕3
= π ξ0 2 
8 2
Thus we can find ξ𝜉0 . Solving for ξ𝜉0 gives
2 × 21/∕4
ξ0 = V3/∕8 (51)
3 π3/∕8
Hence the position of the wetting line R(t) in terms of the original variables is
2 × 21/∕4
R (t) = ξ0 τ1/∕8 = V3/∕8 τ1/∕8
3/∕8
3 π
1/∕4
2×2 ρ g V3 1/∕8
=  t
3μ (52)
3 π3/∕8
8 SpreadingAxisymetricDrop.nb

ρ g V3 1/∕8
= 0.893914  t

We can also determine the shape of the spreading drop in terms of original variables
Substituting Eq. (49) into Eq. (34) gives
h (τ, ξ) = τα f (ξ)

3 1/∕3
1/∕3
= τ-−1/∕4 ξ20 -− ξ2 
16

3 1/∕3 ξ 2 1/∕3
= τ-−1/∕4 ξ2/∕3
0 1 -− 
16 ξ0

1/∕3 2 1/∕3 (53)


3 r
= τ-−1/∕4 ξ2/∕3
0 1 -− 
16 r0

3μV 1/∕4 3 1/∕3 r 2 1/∕3


= (0.893914)2/∕3 1 -− 
ρgt 16 r0

3μV 1/∕4 r 2 1/∕3


= 0.531126 1 -− 
ρgt r0
which agrees with Middleman’s formula.

References
◼ T. D. Blake, A. Clarke, J. De Coninck and M. J. de Ruijter, “Contact angle relaxation during droplet
spreading: Comparison between molecular kinetic theory and moleciular dynamics.”, Langmuir,
13,1997 pp. 2164-2166.
◼ J. Lopez, C. A. Miller and E. Ruckenstein, “Spreading kinetics of liquid drops on solids”, J. Colloid
Interface Sci., 56, pp 460-468, 1976.
◼ H. Hu and R. G. Larson, “Analysis of the microfluid flow in an evaporating sessile droplet”, Langmuir,
21, pp 3963-3971, 2005.
◼ S.N. Reznik and A.L. Yarin, “Spreading of an axisymmetric viscous drop due to gravity and capillarity
on a dry horizontal wall”, Int. J. Multiphase Flow, 28, pp 1437-1457, 2002

Appendix

Curvature Term
(52)
For an axisymmetric drop the twice the mean curvature of the interface can be expressed as
SpreadingAxisymetricDrop.nb 9

1 1
2H= + , where r1 , r2 are the principal curvatures of the interface
r1 r2
where r1 and r2 are the principal curvatures of the interface. We can express the height of the interface
by h=h(r) . Then the curvature becomes
ⅆ2 h  dr2 ⅆh /∕ ⅆr
2H= +
3/∕2 1/∕2
1 + (ⅆh /∕ ⅆr)2  r 1 + (ⅆh /∕ ⅆr)2 
where r and z=h are the cylindrical coordinates of the spreading drop. If the drop curvature is small, we
can neglect the terms in the denominator, i.e. (∂𝜕h/∕∂𝜕r)2 <<1, to get
ⅆ2 h 1 ⅆh 1 ⅆ ⅆh
2H≈ + = r
ⅆr2 r ⅆr r dr ⅆr

You might also like