Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Computers & Fluids 82 (2013) 110–121

Contents lists available at SciVerse ScienceDirect

Computers & Fluids


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 / c o m p fl u i d

Some experiences with the numerical simulation of Newtonian


and Bingham fluids in dip coating
Abdelkader Filali a, Lyes Khezzar a,⇑, Evan Mitsoulis b
a
Department of Mechanical Engineering, The Petroleum Institute, P.O. Box 2533, Abu Dhabi, United Arab Emirates
b
School of Mining Engineering & Metallurgy, National Technical University of Athens, Zografou 157 80, Athens, Greece

a r t i c l e i n f o a b s t r a c t

Article history: The dip-coating process is simulated numerically for Newtonian and Bingham fluids with a particular
Received 18 September 2012 emphasis on finding the free surface location under different sets of conditions. The main focus of the
Received in revised form 12 April 2013 simulations is the evaluation of the Finite-Volume Method (FVM) combined with the Volume-Of-Fluid
Accepted 22 April 2013
(VOF) technique embedded in the commercial code (Fluent), while some results are also obtained with
Available online 9 May 2013
the Finite-Element Method (FEM) using another commercial code (Polyflow). The objective is to check
how well the results compare with previous results regarding the free surface location for both axisym-
Keywords:
metric and planar geometries, and also provide new results. The numerical results were first validated
Dip coating
Free surfaces
against previous experimental data for the Newtonian limit case. Then numerical results were compared
VOF method favourably with previous simulations available in the literature for both Newtonian and Bingham fluids.
Inertia The influence of the coating fluid properties as well as surface withdrawal speed for Newtonian fluids and
Surface tension of the yield stress for Bingham fluids are discussed. The effect of inertia for the planar case is investigated,
Newtonian fluid and the formation of the cusp and wavy shape of the free surface is analyzed.
Bingham fluid Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction ume-Of-Fluid (VOF) technique of Hirt and Nichols [3], are in prin-
ciple not limited in their range of applicability, and offer a strong
Coating flows are liquid flows forming on solid surfaces. Dip- viable alternative. The present work uses such techniques in the
coating is a process that belongs to the family of self-metered coat- study of dip-coating flows with the objective of finding their limi-
ing processes, wherein a substrate is dipped into a liquid coating tations and of gaining experience in applying them in cases where
solution and then is withdrawn from the solution at a controlled the free surfaces are highly curved.
speed (Fig. 1). The combination of substrate speed, liquid proper- Three non-dimensional numbers emerge to describe this flow.
ties (such as density, viscosity, and surface tension), and the pro- These are:
cess geometry influence the thickness of the coated film. The
produced film has to satisfy thickness and uniformity require- (i) the capillary number, Ca, which represents the ratio of vis-
ments of the final product design. cous to surface tension forces and is defined by:
Dip-coating processes are found in a wide field of practical
lV 0
applications. These coatings are used for decorative, protective Ca ¼ ; ð1Þ
and functional applications that include galvanized steel, magnetic
r
information storage systems and manufacture of semi-conductor where l is the viscosity, r is the surface tension, and V0 is a charac-
components. Although coating processes have been developed teristic speed (in dip coating the substrate speed);
mainly through empirical means, see Ruschak [1] and Weinstein (ii) the Reynolds number, Re, which represents the ratio of iner-
and Ruschak [2], appropriate design and manufacture calls for a tia to viscous forces and is defined by:
detailed understanding, knowledge and ability to predict the coat- qV 0 lc
ing flow process. Computational Fluid Dynamics (CFD) methods, Re ¼ ; ð2Þ
l
based on the solution of the governing conservation equations,
combined with interface-tracking techniques, such as the Vol- where lc represents the characteristic length, defined as lc = (r/qg)1/2,
where g is the acceleration of gravity;
(iii) the fluid property number, P0, which is equal to the square
⇑ Corresponding author. Tel.: +971 50 820 1536. root of the ratio Ca/Re with the capillary rise (r/qg)1/2 used
E-mail address: lkhezzar@pi.ac.ae (L. Khezzar). as the length scale, is defined by:

0045-7930/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.compfluid.2013.04.024
A. Filali et al. / Computers & Fluids 82 (2013) 110–121 111

method was based on a prediction/projection fixed-point iterative


scheme and a combination of an original nodal displacement
scheme in combination with B-spline smoothing and remeshing
to update the free surface. Hurez and Tanguy [16] carried out a
FEM analysis using Newtonian and non-Newtonian Bingham fluids
with an iterative method based on the Arbitrary Lagrangian–Eule-
rian (ALE) approach in conjunction with mesh smoothing. It was
shown that the influence of the viscosity and the withdrawal
velocity must be studied separately, and that the capillary number
is not sufficient to describe the meniscus flow dynamics, so that as
mentioned above, a fluid property number P0 is also needed.
Yamamoto et al. [17] used FEM to simulate transient viscoelastic
dip-coating flows around a rod with three Phan–Thien–Tanner
(PTT) fluids that differ in the stretch-thickening property. Recently,
Jin et al. [18], using the FIDAP software (Ansys, Inc.), confirmed the
asymptotic Newtonian value of the flow rate of 0.582 for large cap-
illary numbers (Ca ? 1) and observed wavy wiggles on the free
Fig. 1. Schematic representation of the dip-coating problem with reference axes surface for large values of capillary number (Ca  3.5) with the
and boundary conditions. property number P0 = 0.83. Jenny and Souhar [19] used FEM with
 1=4 rffiffiffiffiffiffi a semi-implicit front-tracking method for the moving free surface
g Ca and identified steady and wavy regimes based on the property P0
P0 ¼ l ¼ : ð3Þ
qr3 Re and capillary numbers Ca for high Reynolds number Re > 1. It
was concluded that beyond a critical Re  5, the flow becomes
It is to be noted that the fluid property number is related to the wavy producing fluctuations of the order of 10% of the mean value
Morton number, Mo ¼ P40 . of the film thickness.
The dip-coating flow is usually laminar in nature, initially tran- It can be said that on the experimental side, most of the previ-
sient and tends towards a steady state, and has a free surface the ous studies were based on flow-visualization techniques and con-
dynamics and final shape of which need to be predicted accurately, sisted of observing and measuring the free-surface conditions.
first during the transient and then steady-state periods. The loca- These efforts, successful as they were, were limited in the amount
tion of the free surface and the fluid rheology in the case of non- of quantitative information that can be obtained. On the other
Newtonian fluids presents particular challenges to the numerical hand, CFD as a simulation tool allows access to useful quantitative
simulation of such flows. information on such flows, which would be difficult to obtain using
Following the initial pioneering work of Landau and Levich [4] experimental tools. It transpires also from the above literature re-
and Derjaguin [5], several experimental investigations were car- view that most numerical studies used FEM with remeshing tech-
ried out, such as by Lee and Tallmadge [6], Tallmadge and Soroka niques. Present-day numerical methods based on the Finite
[7], Groenveld [8], and Spiers et al. [9]. Kizito et al. [10] carried Volume Method (FVM) and combined with an interface tracking
out experiments in a broad parametric range, including extended technique, such as VOF, are capable of simulating dip-coating flows
Ca and Re numbers. Using laser-induced fluorescence, they identi- and can be used for analysis and design purposes.
fied the phenomena leading to the formation of an asymptotic In the present work, an analysis of the dip-coating process for a
meniscus profile, which eventually develops a cusp at the interface vertical plane and a cylindrical rod is carried out using FVM
when inertia forces are important (Re > 7). Based on the property embedded in ANSYS – FLUENTÒ code version 12.1 [20]. Both New-
number, P0, two phenomena of free coating are identified. When tonian and non-Newtonian Bingham fluids are used. The location
P0 is larger than about 0.5, the non-dimensional final film thickness of the free surface at equilibrium, in the presence of surface tension
becomes constant beyond Ca  1. For Ca < 1, the data shows that and external gravity force, is predicted starting from an initial li-
the effect of Re is negligible and the film thickness depends only quid bath at rest with a horizontal free surface, using VOF for inter-
on the capillary number. However for Ca > 1, the viscous force pre- face capturing. The present study provides an evaluation of the
dominates, and surface tension has a negligible influence in that VOF algorithm and FVM in predicting the behavioral details of such
situation. When P0 is less than about 0.1, the non-dimensional final flows. The numerical work is contrasted with the results obtained
film thickness depends on Ca and Re but it becomes constant be- in the works by Kizito et al. [10] and Hurez and Tanguy [16]. Par-
yond a Weber number (We = Ca Re) of about 0.2. In that case Re ticular attention is given to the existence of cusps on the free sur-
is important, and the surface tension force becomes negligible face under certain flow conditions and the resulting flow field
compared with the inertia force. In both cases, the non-dimen- structure for dip coating in planar geometries.
sional final film thickness becomes constant as the effect of surface
tension on the meniscus becomes relatively unimportant. 2. Mathematical modeling
Numerical simulations based on the Finite Element Method
(FEM) have been carried out by Tanguy et al. [11,12], who modeled Since the focus of this study is an evaluation of the technique
the dip-coating process for Newtonian and non-Newtonian fluids. based on the FVM-VOF combination, its mathematical model is de-
Numerical results were presented for both flat-plate coating (pla- scribed in some detail, while a short description of the FEM mod-
nar geometry) and wire coating (axisymmetric geometry), and eling approach is also given for completion.
compared successfully with experimental data by Lee and Tall-
madge [13]. Most previous numerical works used FEM. Tanguy 2.1. Governing equations
et al. [12] used the Augmented Lagrangian Method developed by
Fortin and Glowinski [14]. An iterative solution based on FEM The field conservation equations for mass and momentum for
was used by Reglat et al. [15] to predict the free-surface flow dis- transient, incompressible, laminar flow are given below and are
placement of a Newtonian fluid by a cylindrical substrate. The shared by the liquid and gas phases:
112 A. Filali et al. / Computers & Fluids 82 (2013) 110–121

r  u ¼ 0; ð4Þ and velocity is achieved using the well-known SIMPLE (Semi-Im-


plicit Method for Pressure-Linked Equations) algorithm [24]. The
Du convergence criteria for the inner iteration at each time step were
q ¼ rp þ r  s þ qg þ F S ; ð5Þ
Dt set to 106 for all the relative (scaled) residuals.
where u represents the velocity vector, p the pressure, s the extra In transient dip-coating fluid motion, two characteristic times
stress tensor, and Fs is the surface tension force (see Brackbill can be identified, a characteristic time scale associated to the first
2
et al. [21]), which will be discussed later. Stokes problem [25] can be defined as tSt  qlc =g where lc repre-
For a generalized Newtonian fluid, the rheological constitutive sents the characteristic length scale defined earlier (Eq. (2)). The
equation that relates the extra stress tensor to the rate-of-strain second one is associated with the stretching or formation of new
tensor is given by interface in the meniscus region, defined as tSt  lc/V0. In the pres-
ent work and for low capillary numbers (Ca < 0.5), the smallest
s ¼ gc_ ; ð6Þ characteristic time scale is the one calculated from the first Stokes
2
where g is the apparent viscosity and c_ represents the rate-of-strain problem tSt  qlc =g. Hence, a minimum non-dimensional time step
4
tensor given by. of 10 , defined by the ratio of the real time step to the character-
istic time scale, is used to avoid the appearance of unphysical wig-
c_ ¼ ru þ ruT ; ð7Þ gles on the free surface. For relatively high capillary numbers
(Ca > 0.5) and to avoid divergence, a non-dimensional time step
where ru is the velocity-gradient tensor and the superscript T is
equal to 104 is initially used and then as the film is created, it is
the transpose of a tensor. The apparent viscosity g is a function of
increased gradually to a maximum value of 102 to reduce the
the magnitude of the rate-of-strain tensor
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi overall calculation time. In the present simulations, 95% of the cell
1 Courant numbers (defined by: Cr = Dt/DxcellVfluid) was below 2.
c_ ¼ jc_ j ¼ ðc_ : c_ Þ; ð8Þ Solving the transient governing equation to obtain a steady-state
2
solution of the flow at very low capillary numbers takes more time
When the apparent viscosity is a constant g = l, then Eq. (6)
than at a high capillary numbers (which corresponds to negligible
represents the Newtonian fluid.
surface tension effects, hence a merely viscous flow). At very low
For viscoplastic non-Newtonian fluids having a yield stress s0,
capillary numbers, one simulation may take up to 8 days of com-
the rheological constitutive equation is described by the Her-
puting time.
schel–Bulkley model [20], which is a generalization of the Bingham
The FEM technique uses a remeshing algorithm to track the
model and combines the effects of the yield stress and the power
fluid interface. Numerically, the interpolation schemes of quadratic
law. For stresses below a yield stress s0, the material behaves like
co-ordinates, quadratic velocity and linear pressure were em-
a rigid solid. As the stresses increase beyond the yield stress
ployed, and a Picard iteration scheme was used to improve conver-
threshold, s0, the material behaves as a fluid described by a power
gence [26]. For the FEM computations using POLYFLOWÒ, all cases
law. The Herschel–Bulkley model is characterized by three param-
converged. The convergence criterion, based on the relative differ-
eters: yield stress, consistency index K, and power-law index n.
ence between successive updates used in our simulations, was
The ideal Herschel–Bulkley model is discontinuous at the yield-
105. For each type of field, the modification at every node be-
stress point and thus very difficult for computations. This difficulty
tween two successive iterations was compared to the maximum
is circumvented by using the bi-viscosity model, introduced by
value of the field at the current iteration. It is to be noted that
Tanner and Milthorpe [22], which makes use of a critical shear rate,
the FEM technique is rather expensive in terms of memory and
c_ c , below which the material becomes very viscous, approaching CPU time requirements, in contrast to the FVM-VOF combination.
the behavior of a rigid solid. Then, the modified Herschel–Bulkley
Although the remeshing technique of FEM is intrinsically more
model obeys Eq. (6) for fluids with the apparent viscosity given
accurate than the VOF method for modest mesh sizes, this advan-
by the following relation [23]:
tage is offset by using sufficiently refined meshes with the VOF.
 n1
s c_
g ¼ _0 þ K _ ; for c_ > c_c ð9aÞ 2.3. Free surface treatment and interface location
c cc
  The simulation of dip coating consists of finding the free surface
ð2  c_ =c_c Þ c_
g ¼ s0 þ K ð2  nÞ þ ðn  1Þ ; for c_ < c_c ð9bÞ shape and location from an arbitrary initial position. The FVM in
c_c c_c
FLUENTÒ uses the VOF algorithm of Hirt and Nichols [3] to capture
As mentioned above, the critical shear rate marks the transition the air–liquid interface. This method has been shown to be more
from solid-like (unyielded) to liquid-like (yielded) behavior and flexible and efficient than other methods for tracking liquid–gas
takes a very small value (equal to 105 in the present work), while interfaces see, e.g. [27,28]. This formulation uses a single phase
c_ is the magnitude of the rate-of-strain tensor given by Eq. (8). In (one fluid) set of conservation Eqs. (4) and (5) and constitutive
the present work, we restrict our simulations to the case of Bing- Eqs. (6) and (9), and differences in material properties and surface
ham fluids for which n = 1 and K = l. tension are accounted for by solving a convection equation for the
volume fraction of the secondary phase aw, assumed to be the li-
2.2. Method of Solution and numerical schemes quid in the present work. Assuming incompressible flow, this
equation, in the absence of mass transfer, takes the following form:
The conservation equations are integrated numerically using
Daw @ðaw Þ
the FVM embedded in ANSYS – FLUENTÒ code version 12.1 [23] ¼ þ u  raw ¼ 0 ð10Þ
in combination with the VOF model to predict the liquid–gas inter- Dt @t
face. The fluid is considered to be a generalized Newtonian fluid The volume fraction for the air aa, the primary phase; will be
obeying Eq. (6) either with a constant viscosity (g = l, Newtonian computed using the following constraint:
fluid) or with the Bingham model relationships of Eq. (9) with
aw þ aa ¼ 1 ð11Þ
n = 1. The first order implicit time integration scheme and the third
order QUICK interpolation scheme is used for both, the momentum In addition to the solution of Eq. (10), the VOF uses an accurate
and the volume-fraction equations, and coupling of the pressure algorithm for advecting the volume fraction so as to maintain con-
A. Filali et al. / Computers & Fluids 82 (2013) 110–121 113

servation of mass. In this work, the geometric reconstruction In order to maintain a near constant bath level, the domain con-
scheme is used to represent the interface between the fluids using tains a small inlet flow to compensate for the liquid withdrawal
a piecewise linear interface construction belonging to the family of from the domain. The flow rate at the inlet boundary condition
methods introduced by Youngs et al. [29]. The resulting velocity was estimated from a relationship provided by Wilson [30].
field from Eqs. (4) and (5) is shared among the phases. The volume Although this relationship is limited to Newtonian fluids it was
surface tension force FS depends on the interface topology. It is also used for the Bingham fluid cases. The variation of the liquid
modeled using the Continuum Surface Force (CSF) method of bath level was monitored during the calculations. The bath level
Brackbill et al. [21], thus: was tracked at every time step during the simulation, and the re-
sults showed a difference of less than 2% between the initial and
qjraw
FS ¼ r qw þqa ð12Þ final liquid bath levels with no impact on the flow structure in
2 the bath. It was concluded that this approach of imposing the flow
rate had no influence on the present results.
where qa and qw are respectively the air and the liquid densities
The dimensions of both the axisymmetric and planar geometry
and j is the curvature defined in terms of the unit normal to the
correspond to an experimental setup on which data for the coating
interface n, which in turn is calculated using the volume fraction
process were obtained. For the axisymmetric geometry the same
and its gradients:
configuration and dimensions of Hurez and Tanguy have been con-
 
raw sidered [16] for the FEM calculations, while an extended domain
j¼rn¼r ð13Þ
kraw k has been used for the FVM calculations. For the planar geometry,
the dimensions were taken from the experimental setup of the
In the one-fluid model formulation volume-fraction-averaged
coating applicator described by Kizito et al. [10]. Due to the sym-
properties are calculated using the following relation:
metry, only half of the geometry is investigated. It consists of a ver-
/ ¼ aw /w þ ð1  aw Þ/a ð14Þ tical flat plate being withdrawn with constant speed V0, from a
liquid bath of effectively infinite depth and semi-infinite horizontal
where / can represent density or viscosity. dimension. Thus, the side wall and the outlet boundary conditions
The film thickness h0 at exit is calculated at every time step, and have a negligible effect on the film thickness values and the free
the final free surface location is reached when h0 become constant. surface shape. As discussed by Kizito et al. [10], the dimensions
It is well known that the VOF method [3] produces a diffused inter- of the applicator depth considered has negligible influence on the
face zone corresponding to artificial mixtures created by numerical data obtained in their work [10].
diffusion. To reduce the numerical diffusion zone and improve on
the accuracy of the final film thickness calculations, a well refined
3. Results and discussion
mesh, in particular at the interface and near the moving plate
where the film is developing, has been used. 20 grid cells were
3.1. Mesh refinement studies
used across the film thickness in the axisymmetric geometry
where Ca was relatively large. For the planar case as many as 40
Before we embark on the effect of the different parameters, it is
grid cells were used across the film thickness.
instructive to present a validation of previous results for the axi-
The diffusion zone at the interface is found to be very small with
symmetric [16] and planar case [10]. The fluid properties are listed
the meshes used. At exit, its extent was less than 6% of the final
in Tables 1 and 2.
film thickness. In the VOF results, the location of the free surface
For the axisymmetric case, Fig. 2 shows the calculation domain
is hence found by creating an iso-surface where the liquid vol-
used with the mesh and boundary conditions. The fluid domain
ume-fraction is equal to 1. The position of this iso-surface is then
considered for the FVM calculations has a width of 0.047 (m) and
extracted easily. The implication is that the position of the inter-
a depth of 0.08 (m), whereas for the FEM calculations the same
face is known with an error of less 6%.
dimensions of the fluid domain available in [16] were used with
0.047 (m) width and 0.063 (m) depth, starting from a guessed free
2.4. Boundary and initial conditions surface shape with an initial bath level at z = z0 = 0.048 (m). A
structured mesh consisting of non-uniform, rectangular cells, with
Two geometrical configurations have been used in this study; refinement near the regions of greatest interest, i.e., near the verti-
the first one concerns dip coating of a rod in an axisymmetric cal-wall regions, the free surface, and the expected vertical film re-
geometry (r, z) [16], and the second one a planar geometry (x, y) gion was used.
of dip coating on an infinite plane [10]. As already mentioned, Grid sensitivity tests were conducted with different numbers of
the solution domain and boundary conditions for the axisymmetric cells to calculate the location of the free surface for the case in the
geometry are shown in Fig. 1. literature [16] with Ca = 6.03, using Fluid N1. For the FVM-VOF
A no-slip condition is adopted on the two walls and the moving method, three mesh sizes were considered: M1 with 18,000 cells,
rod/plate. Thus the substrate moves at v = V0, while the non-mov- M2 with 34,875 cells and M3 with 57,000 cells, respectively. The
ing walls have zero velocities. The upper edge of the air-filled do-
main was modeled as a zero gauge pressure boundary condition.
In the case of the FVM-VOF combination, the free surface profile Table 1
is not known a priori, and the problem is solved as a transient one Physical properties of fluids [16].

starting with an initial horizontal free liquid surface. The position Parameters Newtonian Bingham
of the free surface is tracked with time until a steady-state condi- Fluid N1 Fluid N2 Fluid B1 Fluid B2
tion is reached when the film thickness reaches a constant final va-
Viscosity, l (Pa s) 2.9 100 2.9 2.9
lue. Hence, initially, at t = 0, the liquid is at rest and the Yield stress, s0 (Pa)   1.0 4.0
computational domain is filled with the air–liquid mixture with Density, q (kg/m3) 910 910 910 910
the liquid level set at a horizontal level z = z0. In contrast, the Surface tension, r (N/m) 0.0359 0.0359 0.0359 0.0359
FEM calculations use a remeshing method to track the free-liquid Substrate speed, V0 (m/s)
V0 = 0.0241 (m/s) Ca = 1.95 Ca = 67.13
interface, starting with a necessary ‘‘good’’ initial guess of the final
V0 = 0.0747 (m/s) Ca = 6.03
free surface position.
114 A. Filali et al. / Computers & Fluids 82 (2013) 110–121

Table 2
Physical properties of fluids for the planar geometry [10].

Fluid Density, q (kg m3) Surface tension, r (N m1) Viscosity, l (Pa s) Fluid property, P0
5 cs Silicone oil 916 0.0197 0.00458 0.03
20 cs Silicone oil 949 0.0206 0.01898 0.1
100 cs Silicone oil 964 0.0209 0.09640 0.56
200 cs Silicone oil 967 0.021 0.19340 1.1
1000 cs Silicone oil 970 0.0212 0.97000 5.5

Pressure outlet

Initial bath level

Wall
Axis of Moving
symmetry rod / plate

Wall Fig. 4. Grid sensitivity test with the FEM-remeshing method. Free surface profile
Fluid for a Newtonian fluid for Ca = 6.03 (axisymmetric case).

Fig. 2. Finite volume mesh with reference axes and boundaries.

meshes M2 and M5 are used in all simulations to optimize the


computation memory and time.
For the planar case, a fluid domain of 0.15 (m) width and 0.35
(m) depth is used, and the flow configuration is the one docu-
mented experimentally in [10]. For very low capillary numbers in
the range of 0.01 < Ca < 0.1, the exit film thickness h0 is very thin
requiring a fine mesh along the interface and solid boundaries.

Fig. 3. Grid sensitivity test with the FVM-VOF numerical method. Free surface
profile for a Newtonian fluid for Ca = 6.03 (axisymmetric case).

location of the free surface is plotted for the three different meshes
in Fig. 3. For the FEM-remeshing method, three mesh sizes were
considered: M4 with 2425 cells, M5 with 4850 cells and M6 with
7300 cells, respectively. The location of the free surface is plotted
for the three different meshes in Fig. 4. As the difference between Fig. 5. Grid sensitivity test for the 2-D planar case using the FVM-VOF numerical
M2, M3 and M5, M6 respectively for both methods, is insignificant; method. Free surface profile for a Newtonian fluid for P0 = 1.1 and Ca = 1.
A. Filali et al. / Computers & Fluids 82 (2013) 110–121 115

a b

Fig. 6. Variation of the free surface shape with capillary number Ca for Newtonian fluids: (a) Ca = 1.95, (b) Ca = 6.03 and (c) Ca = 67.13 (axisymmetric case). Dimensional
results with different methods.

Numerical tests were conducted for different Newtonian fluids calculated free-surface shape with the experimental data of [16].
with their physical properties listed in Table 2 [10]. Good agreement is shown between the present work and the
Grid sensitivity tests were conducted with different numbers of experimental data of Hurez and Tanguy [16] for Ca = 6.03 with a
cells to calculate the location of the free surface for a Newtonian difference of 5% for the final film thickness and a maximum of
fluid with P0 = 1.1 and Ca = 1. Four mesh sizes were considered: 8% in the dynamic meniscus region.
M7 with 216,550 cells, M8 with 313,950 cells, M9 with 414,050 The comparison with numerical results obtained by Hurez and
cells and M10 with 673,200. The location of the free surface is plot- Tanguy [16] for the shape of the free surface for different capillary
ted for the four different meshes in Fig. 5. Since for low capillary numbers Ca can also be seen in Fig. 6a–c respectively. For the VOF
numbers the film thickness is very small and needs more refine- model, the difference between the present results and those of
ment near the walls and the interface zones, mesh M6 is used in Hurez and Tanguy [16] is about 5% for the final film thickness
all simulations. and a maximum of 7% in the dynamic meniscus region for
Ca = 1.95, see Fig. 6a, and about 7% for the final film thickness
3.2. Axisymmetric dip coating – Newtonian fluids and a maximum of 10% in the dynamic meniscus region for
Ca = 6.03, see Fig. 6b. For the FEM the difference is about 2% for
Numerical results are validated with experimental data on fluid the final film thickness and 9% in the dynamic meniscus region
N1 (Table 1) used in dip coating experiments [16] with the FVM- for Ca = 1.95, and about 4% for the final film thickness and 7% in
VOF combination. Additional numerical results based on the the dynamic meniscus region for Ca = 6.03.
FEM-remeshing technique using Polyflow software [26] are pre- However, Fig. 6c for high Ca = 67.13 shows a clear disagreement
sented for comparison between both approaches in terms of free- when compared with the Hurez and Tanguy results [16]. By taking
surface profiles, final film thickness and streamline patterns. a closer look it is clear that the results of the present work obtained
In the FEM-remeshing method, the initial free surface shape is with the two different numerical methods (FVM and FEM) appear
guessed, which is not the case for the FVM-VOF method where to be closer in terms of final film thickness, location of stagnation
the initial free surface corresponds to the position of the fluid at point, and free surface. The disagreement is confined to the menis-
rest, i.e., z = z0 = constant. Fig. 6b illustrates the comparison of the cus region and is on the order of 30%. There are potentially two
116 A. Filali et al. / Computers & Fluids 82 (2013) 110–121

Newtonian Fluid a
Stagnation
Hurez and Tanguy [16]
FVA - VOF point
FEM-Remeshing
Experimental [16]
Line for visual aid

Liquid VFr = 1
Air VFr = 0
z (m)

Fig. 7. Variation of the dimensional final film thickness h0 with capillary number Ca
for Newtonian fluids (axisymmetric case). Results with different methods.

main reasons for this discrepancy at this high Ca number. First the r (m)
number of elements used in [16] was rather modest (around 200)
compared to the present mesh sizes. When Ca is high, the displace- b
ment of the free surface during the calculations with the reme- Stagnation
point
shing technique is important, and using a relatively coarse mesh
leads to the distortion of the cells and eventually to divergence
as we have experienced it. A high Ca leads to a relatively larger film
thickness, and this required us to extend the meshed domain in the
vertical direction, thus providing an exit boundary condition far
away from the liquid bath free surface. It appears that the meshed z (m)
domain used by Hurez and Tanguy [16] for Ca = 67.13 was not long
enough in the vertical direction to have an exit boundary far re-
moved from the liquid bath surface to allow a good resolution of
the final film thickness. This could be the second reason to explain
the discrepancy between the present results and those of [16] as
seen in Fig. 6c.
The variation of the final film thickness h0 with capillary num-
ber Ca in a system of logarithmic coordinates is presented in Fig. 6.
The numerical results for both FVM and FEM techniques are com- r (m)
pared with experimental and numerical data of Hurez and Tanguy
Fig. 8. Streamlines patterns superimposed onto contour plots of the volume
[16]. As shown in Fig. 7, the final film thickness increases linearly fraction of the Newtonian fluid N1 (VFr = 1) for Ca = 6.03: (a) FVM-VOF, (b) FEM-
in log scale with capillary number. For Ca = 1.95 and 6.03, the remeshing technique (axisymmetric case).
FEM results are closer to the numerical and experimental results
of Hurez and Tanguy [16] than the FVM results, whereas for high
Ca (Ca = 67.3), the FVM and FEM techniques agree in terms of the as Bn = s0/(lV0/h0). Fig. 9 presents a comparison between the pres-
final film thickness but disagree with the numerical results of Hur- ent results and numerical results available in the literature [16] for
ez and Tanguy [16] as discussed above. Bn = 0.173 and 0.7, respectively. Sensitivity to the critical shear rate
As an illustration of the behavior of the free surface in dip coat- has been tested by using a smaller critical shear rate c_ c ¼ 108 s1
ing, streamline patterns superimposed onto contour plots of the than the value of c_ c ¼ 105 s1 , which was the standard value used
volume fraction are shown in Figs. 7a and 8b for Ca = 6.03 for the in all of our computations. It was found that there was no discern-
FVM-VOF and the FEM-remeshing technique, respectively. The fi- ible effect on the final film thickness and the position of the stag-
nal meniscus shape and flow field show the sharp separation of nation point, except for a small difference of less than 1% in the
the flow into two regions: in the upper part, the fluid is entrained dynamic region of the free surface.
along the web while in the lower part the fluid returns to the res- For Bn = 0.173 in Fig. 9a, results of the present work obtained
ervoir and generates relatively intense recirculation. with the two different numerical methods (FVM and FEM) are very
As a comparison between the two methods for Ca = 6.03 in similar to each other. There are, however, differences with the
Fig. 8a and b, respectively, and as shown previously in Fig. 6b, numerical results of [16]. Thus, although a fairly good agreement
the results are close in terms of the free surface shape, streamline on the location of the free-surface shape in the upper region of
patterns, and the position of the stagnation point in the x- and y- the meniscus in which the film thickness becomes constant exists,
directions, with a difference less than 3%. There is no discernible and also close to the surface of the bath with a difference of 6%, a
difference (less than 0.5%) in the location of the eye of the flow relatively large difference of about 20% exists in the meniscus dy-
recirculation zone inside the bath between the FVM and FEM namic region in which the film thickness varies with the height
calculations. above the bath level and in which viscous, gravity, and surface ten-
sion have all to be taken into account. Results for Bn = 0.7 in Fig. 9b
show that the free surface shape using the FVM-VOF technique
3.3. Axisymmetric dip coating – Bingham fluids agree with the numerical results obtained by Hurez and Tanguy
[16], while for the FEM-remeshing technique the results show a
For the Bingham fluid, numerical simulations are conducted difference of 15% in the dynamic region of the free surface but
using two fluids B1 and B2 for different values of the yield stress agree well with the FVM-VOF and Hurez and Tanguy [16] results
s0 and/or the dimensionless Bingham number, which is defined in terms of the final film thickness.
A. Filali et al. / Computers & Fluids 82 (2013) 110–121 117

Fig. 10 represents the variation of the free-surface shape for the


a
Bingham Fluid (Bn = 0.173)
two Bingham fluids B1 and B2 for different values of the yield
stress s0 or Bn as well as the Newtonian limit case (Bn = 0). It can
Hurez and Tanguy [16]
be noticed from Fig. 10 that entrainment of the free surface shape
FVA - VOF
FEM-Remeshing for Bingham fluids, Bn = 0.173 and 0.7 respectively, exists only in
the dynamic region. In contrast to the Newtonian fluid, the entrain-
ment of the free surface for Bn = 0.173 is around 4% and for Bn = 0.7
is 6%.
Of interest is the extent and shape of the yielded/unyielded (un-
shaded/shaded) regions separated by the contour |s| = s0, where |s|
is the magnitude of the stress tensor. When the Bingham number is
successively increased from Bn = 0.173 to Bn = 0.7 to Bn = 10, the
results are given in Fig. 11 [31]. It is shown that increasing the yield
stress affects the free-surface shape and tends to increase the
entrainment of the coating liquid. The comparison among these
cases shows that increasing the Bn number changes the shape
and extent of the yielded/unyielded regions. In particular increas-
ing Bn (a dimensionless yield stress) leads to a progressive increase
of the unyielded (shaded) region, as expected since the material
becomes more solid-like [31]. Most of the unyielded regions are
b found away from the meniscus and the free surface, in the upper
Bingham Fluid (Bn = 0.7)
area of the reservoir, where the flow is slow and not changing very
Hurez and Tanguy [16]
much. There are yielded areas in the reservoir near the entry and
FVA - VOF
the walls, where shearing occurs for the lower Bn numbers of
FEM-Remeshin g
Bn = 0.173 and 0.7. On the other hand, near the meniscus and the
free surface for the same Bn numbers, yielding occurs due to curva-
ture and shearing on the moving web. However, at higher Bn num-
bers (Bn P 10) the unyielded region is more important and covers
most of the flow domain, even under the free surface and the
meniscus, because of its small curvature, and a very uniform
plug-like velocity profile (Fig. 11c).
By considering a constant fluid viscosity (K = l, n = 1) and a con-
stant withdrawal velocity of the web, the position of the stagnation
point is not affected, but it is clear that increasing the yield stress
tends to increase entrainment and to influence the velocity field as
shown in Fig. 12. This figure represents the streamline patterns
superimposed onto contour plots of the volume fraction of the
Bingham fluid B2. It is seen that the fluid velocities tend to vanish
in the bath, limiting the zone of influence of the moving wall to the
Fig. 9. Predicted free surface location and results by Hurez and Tanguy [16] for
Bingham fluids: (a) Bn = 0.173, (b) Bn = 0.7 (axisymmetric case). Results with near field, where the fluid has a tendency to ‘‘stick’’ to the web
different methods. causing the effect of ‘‘swelling’’, as it has been mentioned in previ-
ous work [16].

3.4. Planar dip coating – Newtonian fluids

Bingham Fluid (FVM-VOF) We now turn our attention to the case of the 2-D planar geom-
Bn = 0 etry with Newtonian fluids, which was studied experimentally in
Bn = 0.173 [10]. Fig. 13 represents the variation of the dimensionless thickness
2
Bn = 0.7 T0 ½T 0 ¼ ðqgh0 =lV 0 Þ1=2  as a function of Ca for several Newtonian
fluids with different Property numbers P0. The comparison of the
Stagnation point
present results with experimental data [6,10] displays a good
agreement with a maximum difference of 12%. For Ca < 1, the
dimensionless thickness T0 increases steeply with increasing Ca,
and the effect of Re (or inertia forces) becomes negligible. For P0
values of 0.03 and 0.1, the steep increase of T0 with Ca is captured
correctly together with a slight overshoot beyond the asymptotic
value. The asymptotic value is reached earlier as P0 decreases.
The correlation developed by Wilson [30] valid for very low values
of Ca  1 is also represented, and it is interesting to see that only
partial agreement with it exists for experimental values of P0 = 0.1.
When P0 = 0.03, both experimental and numerical results exhibit
higher values for T0 than the Wilson line and a steeper variation
with P0. The numerical results of Jenny and Souhar [19] are how-
ever closer to the Wilson line albeit at low values of Ca. But overall
Fig. 10. Variation of the free surface shape with the Bingham number Bn for the departure from the Wilson line was also observed by Jenny and
Bingham fluids using the FVM-VOF numerical method (axisymmetric). Souhar [19]. For P0 = 5.5 experiments and calculations display a
118 A. Filali et al. / Computers & Fluids 82 (2013) 110–121

Fig. 11. Contour plots of the yielded (white) and unyielded (black) regions separated by the line (|s| = s0) for (a) Bingham fluid B1 with Bn = 0.173, (b) Bingham fluid B2 with
Bn = 0.7 and (c) for Bn = 10 (axisymmetric case).

2
Fig. 13. Comparison of the non-dimensional final thickness T0 ½T 0 ¼ ðqgh0 =lV 0 Þ1=2 
b as a function of Ca for different Newtonian fluids with experimental [10] and
numerical [19] and Wilson‘s model [23].

underpredicting the asymptote. For P0 equal to 1.1 the calculations


follow closely the experimental values up to Ca = 0.5, beyond
which the calculation reveal a monotonic increase towards the
asymptotic value in contrast to the experiments which show a sud-
den earlier jump, albeit with some scatter. Thus when P0 > 0.5, T0
becomes constant beyond Ca = 1, which means that the surface
tension forces are no longer influential in this case. The present cal-
culations, shown in Fig. 13, indicate an asymptotic value of
T0 = 0.664 ± 0.03 as Ca is increased, which is in agreement with
the asymptotic value of T0 = 0.67 ± 0.02 obtained by Kizito et al.
[10] and the value of T0 = 0.68 obtained by Ruschak [1].
As a demonstration and validation of the present numerical re-
sults, Fig. 14 represents the comparison of the interfacial profile
between the present work and experimental results by Kizito
Fig. 12. Streamline patterns superimposed onto contour plots of the volume et al. [10] as well as experimental results by Lee and Tallmadge
fraction of the Bingham fluid B2 (VFr = 1) obtained with different methods: (a) FVM- [6] for P0 = 1.1 and three different values of Ca and Re. For
VOF, (b) FEM-remeshing technique (axisymmetric case).
Ca = 0.95, the numerical predictions correlate very well with the
experimental values of [6,10], with the values of [10] being closer
to the calculations. For Ca = 2 and 2.5 it is clear that both calcula-
good agreement showing an increase followed by a levelling-off to- tions and experiments confirm that beyond Ca = 2 surface tension
wards the asymptotic value around Ca = 1 with the calculations forces become negligible when compared with other forces so that
A. Filali et al. / Computers & Fluids 82 (2013) 110–121 119

Fig. 15 compares interfacial profiles between the present


numerical results and the experimental data obtained by Kizito
et al. [10] in the range where inertia forces are important
(We > 2) and the film thickness is in the asymptotic range. In both
cases considered (P0 = 1.1, Ca = 4.45, Re = 7.6) and (P0 = 0.1,
Ca = 0.4, Re = 19.5). In these two cases the interface in the meniscus
exhibits a depression displaying a cusp [10]. Although the differ-
ence between the two experimental interfaces is slight, it is still
observable. The numerically predicted interfaces show a more pro-
nounced depression, and for the case with Re = 19.5 they display an
overshoot beyond the mean horizontal liquid pool line before join-
ing it again, in contrast to the case with Re = 7.9, which shows a
smooth variation of the interface.
In dip coating, a moving plate being withdrawn out of a liquid
entrains a film with a certain thickness and creates a stagnation
point located on the interface of the air–liquid, thus separating
the flow entrained up the plate and the flow back into the bath.
As shown in Figs. 16a and 17a for the velocity magnitude and
Figs. 16c and 17c for the velocity vectors, respectively, below the
stagnation point and for high Re, there exists a locally large accel-
eration of the fluid. As a consequence, the pressure decreases lo-
cally (Figs. 16b and 17b) in the flow direction and the liquid
Fig. 14. Comparison of the free surface location with experimental results of Kizito meniscus dips below the liquid pool line. The interface then be-
et al. [10] and Lee and Tallmadge [6] for P0 = 1.1. Effect of capillary number, Ca
comes eventually horizontal when the pressure recovers as the
(planar case).
meniscus goes deeper into fluid, the fluid particles decelerate and
the meniscus dips back towards the fluid. The decrease and later
the interfacial profile becomes independent of Ca. Looking closely increase of the pressure, as shown in Figs. 16b and 17b, creates a
at the profiles, pronounced discrepancies of the order of 6% be- cusp at the bottom of the depression. This cusp can be conducive
tween calculations and experiments exist in the meniscus region. to air entrainment and hence is not desirable. Fig. 17c reveals com-

3.5. Inertia effects

During high-speed withdrawal of low-viscosity liquids, inertial


Velocity a
magnitude (m/s)
forces become important and influence the meniscus shape near
the liquid pool line. These forces have been neglected in most with-
drawal theories and numerical studies known to the authors
[1,5,7–9]. Relatively few works focused on the effect of inertia
forces in dip coating, e.g., Kizito et al. [10]. In order to gain a better
understanding of the effect of inertia, additional numerical simula-
tions within a range of Reynolds numbers (Re = 7.6 and 19.5) are
conducted and discussed below.

Static pressure
(Pa)
b

Stagnation
point c

z = z0

Fig. 15. Comparison of the interfacial profiles with experimental data of Kizito et al. Fig. 16. (a) Contours of the velocity magnitude |u|, (b) static pressure, and (c)
[10] for high Ca and high Re (planar case). velocity vectors, P0 = 1.1, Ca = 4.45, Re = 7.6 (planar case).
120 A. Filali et al. / Computers & Fluids 82 (2013) 110–121

Velocity
magnitude (m/s)
a

Static pressure
(Pa)

Fig. 19. Location of the wavy and non-wavy state in the log (Ca) and log (1/m).
Stagnation
point
dip coating process. The benefit of one-fluid formulation is in pro-
c viding information on the air–fluid motion near the interface as it
is a natural part of the solution. This can help in the interpretation
and prevention of undesirable interface effects.
Figs. 17a and 18b show the streamline patterns superimposed
z = z0 onto contours of the volume fraction for both cases exhibiting a
cusp, P0 = 1.1, Ca = 4.45, Re = 7.6 and P0 = 0.1, Ca = 0.4, Re = 19.5,
respectively. In both cases Re = 7.6 and 19.5, two different behav-
iors are expected in line with the results by Kizito et al. [10]. In
the first the meniscus profile goes back smoothly after the dip, in
Fig. 17. (a) Contours of the velocity magnitude |u|, (b) static pressure, and (c) contrast with the second case for high Re (Re = 19.5), where the
velocity vectors, P0 = 0.1, Ca = 0.4, Re = 19.5 (planar case). meniscus goes above and below the horizontal pool line with
incipient interface folding before it becomes eventually horizontal.
It should be noticed also that no recirculation zones were predicted
plex air motions adjacent to the free surface. These recirculation
in the hump below the interface as mentioned by Kizito et al. [10],
zones generate shearing on the interface, leading eventually to
and the fluid particles move smoothly through the curved shape of
possible air entrainment in the liquid, which is not desirable in
the free surface as confirmed by the velocity vectors in Figs. 16c,
17c and 18, respectively.
From the different simulations considered, it is found that for a
variety of the parameters examined (property number, capillary
a number and Reynolds number), the shape of the free surface can
be wavy depending on the range of the above parameters. Fig. 19
Stagnation represents the location of the wavy and non-wavy state cases for
point
the different fluids considered in this study. According to [18],
the separation line defined by 1/m  0.12Ca1.35 (m is a fluid prop-
erty number defined in [19] as m = qr3/l4g) is an approximation
of the limit between the wavy and non-wavy states. Clearly and
as expected, the cusp cases for (P0 = 0.1, Ca = 0.4 and P0 = 1.1,
Liquid V Fr = 1 Ca = 4.45) lie within the wavy region according to the separation
b Air V Fr = 0 defined by Jenny and Souhar [19].
Stagnation
point
4. Conclusions

A combined FVM-VOF numerical method was successfully ap-


plied to the computation of dip-coating flows under several dy-
namic conditions. The numerical simulation provides a means to
obtain details and quantitative information on the flow field that
can be difficult if not impossible to obtain using experimental
Fig. 18. Streamline patterns superimposed onto contour plots of the volume
means.
fraction (VFr = 1) using FVM-VOF for both cusp cases: (a) P0 = 1.1, Ca = 4.45, Re = 7.6, Numerical calculations were performed using Newtonian and
and (b) P0 = 0.1, Ca = 0.4, Re = 19.5 (planar case). Bingham fluids for both flat plate coating and wire coating and
A. Filali et al. / Computers & Fluids 82 (2013) 110–121 121

compared successfully with experimental and for some cases with [10] Kizito JP, Kamotani Y, Ostrach S. Experimental free coating flows at high
capillary and reynolds number. Exp Fluids 1999;27:235–43.
numerical data available in the literature.
[11] Tanguy P, Fortin M, Choplin L. Finite element simulation of dip coating, I:
The effect of high inertia forces on the free surface are ade- Newtonian fluids. Int J Num Meth Fluids 1984;4:441–57.
quately captured with a clear exhibit of surface waviness and a [12] Tanguy P, Fortin M, Choplin L. Finite element simulation of dip coating, II: non-
tendency to ingest air for cases displaying a cusp. Calculations also Newtonian fluids. Int J Num Meth Fluids 1984;4:459–75.
[13] Lee CY, Tallmadge JA. Description of meniscus profiles in free coating II –
demonstrate that the recirculation region below the cusp men- analytical expressions. AIChE J 1973;19:403–7.
tioned by Kizito et al. [10] does not exist. However, air recirculation [14] Fortin M, Glowinski R. Méthodes de Lagrangien augmente. Application a la
above the interface in the meniscus region can exist and lead to air résolution numérique de problèmes aux limites. Paris: Dunod-Bordas; 1982.
[15] Reglat O, Labrie R, Tanguy PA. A new free-surface model for the dip coating
ingestion in the wavy regime. process. J Comput Phys 1993;109:238–46.
Overall, the successful completion of various different simula- [16] Hurez P, Tanguy PA. Finite element analysis of dip coating with Bingham
tion tests under different conditions suggests that the VOF ap- fluids. Polym Eng Sci 1990;30:1125–32.
[17] Yamamoto T, Nojima N, Mori N. Numerical analysis of flow behavior in
proach can be useful for coating process design and optimization meniscus region of viscoelastic dip coating flows. J Textile Eng 2005;51:21–8.
studies. [18] Jin B, Acrivos A, Münch A. The drag-out problem in film coating. Phys Fluids
2005;17:103603.
[19] Jenny M, Souhar M. Numerical simulation of a film coating flow at low
Acknowledgement capillary numbers. Comput Fluids 2009;38:1823–32.
[20] Herschel WH, Bulkley R. Consistency measurements of rubber benzene
Financial support for Mr. A. Filali towards his graduate studies solutions. Kolloid-Zeitschrift 1926;39:291–300 [in German].
[21] Brackbill JU, Kothe DB, Zemach CA. Continuum method for modeling surface
from the Petroleum Institute is gratefully acknowledged.
tension. J Comput Phys 1992;100:335–54.
[22] Tanner RI, Milthorpe JF. Numerical simulation of the flow of fluids with yield
References stress. In: Taylor C, Johnson JA, Smith WR, editors. Num meth lam turb flow.
Proc 3rd Int Conf, Seattle. Swansea: Pineridge Press; 1983. p. 680–90.
[1] Ruschak KJ. Coating flow. Annu Rev Fluid Mech 1985;17:65–89. [23] ANSYS FLUENT 12.0/12.1 Documentation; 2009. <www.ansys.com>.
[2] Weinstein SJ, Ruschak KJ. Coating flows. Annu Rev Fluid Mech 2004;36:29–53. [24] Patankar SV, Spalding DB. Calculation procedure for heat, mass and
[3] Hirt CW, Nichols BD. Volume of fluid (VOF) method for the dynamics of free momentum transfer in 3-dimensional parabolic flows. Int J Heat Mass
boundaries. J Comput Phys 1981;39:201–25. Transfer 1972;15:1787–806.
[4] Landau LD, Levich BV. Dragging of a liquid by a moving plate. Acta Physico- [25] Schlichting H. Boundary-layer theory. New York: McGraw-Hill; 1968.
Chim USSR 1942;17:42–4. [26] ANSYS POLYFLOW 12.1 user’s guide; September 2009.
[5] Derjaguin BV. On the thickness of a layer of liquid remaining on the walls of [27] Lakehal D, Meier M, Fulgosi M. Interface tracking towards the direct
vessels after their emptying, and the theory of the application of simulation of heat and mass transfer in multiphase flows. Int J Heat Fluid
photoemulsion after coating on the cine film. Acta Physico-Chim USSR Flow 2002;23:242–57.
1943;20:349–52. [28] Rider WJ, Kothe DB. Reconstructing volume tracking. Technical report LA-UR-
[6] Lee CY, Tallmadge JA. Meniscus shapes in withdrawal of flat sheets from liquid 96-2375. Los Alamos Laboratory Repot; 1996.
baths. Dynamic profile data at low capillary numbers. Ind Eng Chem Fund [29] Youngs DL, Morton KW, Baines ML. Time-dependent multimaterial flow with
1974;13:356–60. large fluid distortion. In: Numerical methods for fluid dynamics. New
[7] Tallmadge JA, Soroka AJ. The additional parameter in withdrawal. Chem Eng York: Academic Press; 1982. p. 273–85.
Sci 1969;24:377–82. [30] Wilson SDR. The drag-out problem in film coating theory. J Eng Math
[8] Groenveld P. High capillary number withdrawal from viscous Newtonian 1982;16:209–21.
liquids by flat plates. Chem Eng Sci 1970;25:33–40. [31] Mitsoulis E. In: Binding DM, Hudson NE, Keunings R, editors. ‘‘Flow of
[9] Spiers RP, Subbaraman CV, Wilkinson WL. Free coating of a Newtonian liquid viscoplastic materials: models and computations’’ invited review in Rheology
onto a vertical surface. Chem Eng Sci 1974;29:389–96. Reviews 2007. London, UK: Brit. Soc. Rheol.; 2007. p. 135–78.

You might also like