Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Capillary imbibition of

royalsocietypublishing.org/journal/rspa non-Newtonian fluids in a


microfluidic channel: analysis
Research
and experiments
Srinivas R. Gorthi1 , Sanjaya Kumar Meher1 , Gautam
Cite this article: Gorthi SR, Meher SK, Biswas
G, Mondal PK. 2020 Capillary imbibition of Biswas2 and Pranab Kumar Mondal1
non-Newtonian fluids in a microfluidic 1 Microfluidics and Microscale Transport Processes Laboratory,
channel: analysis and experiments. Proc. R.
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

Soc. A 476: 20200496. Department of Mechanical Engineering, Indian Institute of


http://dx.doi.org/10.1098/rspa.2020.0496 Technology Guwahati, Guwahati 781039, India
2 Department of Mechanical Engineering, Indian Institute of

Received: 25 June 2020 Technology Kanpur, Kanpur 208016, India


Accepted: 23 September 2020 GB, 0000-0002-8306-0767

We have presented an experimental analysis on the


Subject Areas: investigations of capillary filling dynamics of inelastic
fluid mechanics non-Newtonian fluids in the regime of surface tension
dominated flows. We use the Ostwald–de Waele
Keywords: power-law model to describe the rheology of the
capillary filling, shear-thinning non-Newtonian fluids. Our analysis primarily focuses
non-Newtonian fluids, Washburn regime on the experimental observations and revisits the
theoretical understanding of the capillary dynamics
from the perspective of filling kinematics at the
Author for correspondence: interfacial scale. Notably, theoretical predictions of
Gautam Biswas the filling length into the capillary largely endorse
e-mail: gtm@iitk.ac.in our experimental results. We study the effects of the
shear-thinning nature of the fluid on the underlying
filling phenomenon in the capillary-driven regime
through a quantitative analysis. We further show that
the dynamics of contact line motion in this regime
plays an essential role in advancing the fluid front in
the capillary. Our experimental results on the filling
in a horizontal capillary re-establish the applicability
of the Washburn analysis in predicting the filling
characteristics of non-Newtonian fluids in a vertical
capillary during early stage of filling (Digilov 2008
Langmuir 24, 13 663–13 667 (doi:10.1021/la801807j)).
Finally, through a scaling analysis, we suggest that the
late stage of filling by the shear-thinning fluids closely

follows the variation x ∼ t. Such a regime can be
called the modified Washburn regime (Washburn 1921
Electronic supplementary material is available
Phys. Rev. 17, 273–283 (doi:10.1103/PhysRev.17.273)).
online at https://doi.org/10.6084/m9.figshare.
c.5172223.

2020 The Author(s) Published by the Royal Society. All rights reserved.
1. Introduction 2
Underlying physical issues involved with the transportation of a very small volume in

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


..........................................................
miniaturized devices/systems are fundamentally important in different areas of microfluidics.
The paradigm demands a thorough understanding of the capillary filling phenomenon [1–8]. In
microfluidic confinement, a small volume of liquid moves upon experiencing the thermodynamic
force (capillary force), which is the culmination of the minimization of the interfacial free energy
of the system (fluid–fluid–solid). It is worth mentioning here that the fluid movement in many
lab-on-a-chip (LOC) platforms largely depends on the strength of the capillary force involved
in the process [9]. Accounting for this aspect, researchers have investigated several avenues to
alter the capillarity effect on the underlying transport, like—to name a few—alteration in surface
chemistry, addition of surfactant in the liquid being transported, external field-induced heating
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

and application of external gradients across the interface [10–13].


Since most of the microfluidic devices/systems depend on the passive control mechanism
integrated with the on-chip platform, the net throughput in these systems/devices is primarily
controlled by the modulation of the dynamics of the contact line formed at the fluid–fluid–solid
interface [14–16]. From the perspective of fluid dynamics at the interfacial scale, the motion of
the three-phase contact line plays a crucial role in dictating the frontal dynamics of the fluid–
fluid interface in the pathways. The macroscopic-flow behaviour is intrinsically dependent on the
microscopic parameters governing the contact line region [17]. As a result, filling of fluid into the
capillary (defined as the temporal advancement of the liquid front along the channel length) is
primarily governed by the dynamics of the contact line motion.
Several attempts toward the alteration in the interfacial dynamics of the contact line motion
have been addressed by researchers in the recent past [12,18,19]. Such avenues essentially control
the spatio-temporal movement of the liquid front in the capillary [19,20]. Some of these attempts
include the imposition of physico-chemical patterning on the surface [16,21–24], manipulation
of flow actuation parameters [16,22,25], modulation of surface properties [26] and application of
external fields [11,12,27]. In all the aforementioned cases, alteration in the magnitude of net force
applied at the three-phase contact line brings about a change in the interface motion, leading to
precise control over the filling dynamics. The underlying issues become even more intricate for
the transportation of non-Newtonian fluids, which may primarily be attributed to the rheology
driven modifications in the viscous force acting on the contact line. As such, on account of
this non-trivial effect of the rheology modulated modification of net force acting on the moving
contact line, the filling dynamics of non-Newtonian fluids into the capillary enhance theoretical
understanding and experimental requirements. For example, the rheological behaviour and
associated transport phenomena of many biofluids are of particular significance in biochemistry
research, and hence the imbibition dynamics of non-Newtonian fluids in a capillary is of great
interest. We must stress upon one important point in this context that seminal investigations on
capillary filling, available in the literature, are primarily focused on the external field driven
modulation of the underlying dynamics. An experimental investigation of the autonomous
(without requiring external forcing) filling of non-Newtonian fluid in a microfluidic channel is an
interesting paradigm considering its twofold novel freatures. First, the dynamics of the contact
line motion in the capillary-driven regime plays a delicate role in advancing the fluid front in
the capillary. Secondly, its relevance and usefulness in the design modification of portable LOC
devices typically deployed for clinical diagnostics.
In the following, we experimentally investigate the filling dynamics of inelastic non-
Newtonian fluids in a microfluidic channel over a surface tension dominated regime. We further
calculate the filling rate in the capillary using theoretical approaches consistent with the fluid
kinematics. We compare our theoretical results with those obtained from the experimental
counterpart. As reported in the theoretical analysis, and subsequently validated by the
experimental investigations [9], the Washburn analysis still works for the prediction of the
capillary flow even in a vertical tube during the early stage (time) of filling. The attributable
physical reasoning behind this underlying behaviour is the insignificant effect of gravity during
the initial time instants of the filling. Through the present analysis, attained in particular
3
through experiments in a horizontal capillary (impact of gravity is trivial), we re-establish the
aforementioned reported argument. This finding verifies the applicability as well as the efficacy

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


..........................................................
of the Washburn analysis in predicting the capillary imbibition phenomenon of non-Newtonian
fluids.

2. Experimental details
(a) Description of fluid rheology and parameters
We consider the Ostwald–de Waele power-law model to describe the rheology of the non-
Newtonian fluids [28–30]. Following this model, the apparent viscosity is the coefficient of
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

proportionality between shear stress and the shear rate given by τ = μapp γ̇ , where μapp
represents the apparent viscosity and γ̇ is the shear rate. This apparent viscosity of these groups
of inelastic non-Newtonian fluids μapp can be approximated as m(γ̇ )n−1 [30–32]. It is to be
mentioned that the apparent viscosity either increases or decreases with the shear rate. The
variable m is known as the flow consistency index with units Pa sn . The dimensionless index n
is known as the flow behaviour index. When n is unity, the power-law fluids represented by
the aforementioned constitutive equation reduce to a Newtonian fluid. The values of n > 1 and
n < 1 represent shear-thickening and shear-thinning fluids, respectively. We consider aqueous
carboxy-methyl-cellulose (CMC) solutions, which are widely deployed to replicate/represent non-
Newtonian fluids [24,29], for the current experiments involving shear-thinning fluids (n < 1). For
the present analysis, we consider the following rheological properties, as given in table 1 [33].
For the aqueous solutions (using 0.2% and 0.4% CMC) of interest, the rheology confirms shear-
thinning behaviour [33]. It is worth adding here that the tabulated rheological properties were
experimentally verified by Som [28] and Som & Biswas [34] using capillary tube viscometer.

(b) Sample preparation


We use a microchannel of rectangular cross-section (L × 2h × 2b) fabricated on a polymethyl
methacrylate (PMMA) substrate for the present experiments. We consider a channel height,
2h = 400 µm, having a fixed width (2b) and length (L) as 500 µm and 3 cm, respectively. Figure 1d
shows the image of the fabricated channel used in the present experimental study. The schematic
diagram of the channel is illustrated in figure 1a. We prepare non-Newtonian fluid (shear-
thinning) by mixing cellulose (CMC powder) (w/w) in deionized (DI) water and store the
prepared sample in a magnetic stirrer for proper mixing.

(c) Experimental setup and methodology


We show in figure 1a the schematic diagram of the experimental facility of the present work.
The laboratory facility is shown in figure 1b, which consists of an inverted microscope (Leica–
DMI3000B) with 10× objective magnification, and high-speed camera (Phantom, Miro-LAB320).
The liquid (aqueous solution of CMC) moves through the microchannel, which is fabricated on
PMMA substrates and covered with a clean glass, and placed on the microscope. We use the
microscope for the bright field visualization of the flow; precisely, the displacement process of
the interface. We consider a frame rate of 200 fps (frames per second) to track the interface of the
binary fluid system in the microchannel. The fluid–fluid interface forms between aqueous CMC
solution and the air already present the microchannel. Induction of aqueous cellulose solution
in the channel by the surface tension effect displaces the air in the process. It may be mentioned
here that, owing to the very small Reynolds number, based on dh , hydraulic diameter Redh ∼
10−3 ( 1) pertinent to this analysis, the flow becomes fully developed immediately as it enters
into the channel. The recorded video is analysed using phantom camera control (PCC) software
to determine the advancement of the interface. By tracking the spatio-temporal advancement of
(a) (b)
4
inverted microscope
inverted microscope

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


micro-channel high speed

..........................................................
camera
micro-channel interface
interface

fluid source

monitor
high speed
camera monitor
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

(c) (d)
b

y q y
z h x h
x z
water-CMC solution air

Figure 1. (a) Schematic diagram depicting the experimental facility; (b) laboratory facility showing different components
involved for the experiments; (c) schematic of the problem for the theoretical analysis from filling kinematics. The interfacial
surface tension σ acting along the tangent to the interface makes an angle θ s (static contact angle) with a solid boundary.
Fluid-A (non-Newtonian) moves from the left to right of the channel (in the positive x-direction) and displacing Fluid-B (air).
(d) Image shows the top view of the fabricated microchannel used in the present experiments. (Online version in colour.)

Table 1. Values of variables for aqueous CMC solutions [28,33].

m (Pa sn ) (flow n (flow


% cellulose consistency behaviour
(w/w) index) index)
0.2 0.010 0.90
..........................................................................................................................................................................................................

0.4 0.025 0.82


..........................................................................................................................................................................................................

the interface, we quantify the filling characteristics at pre-determined time intervals. We take the
microchannel for a thorough rinsing with ethanol and Milli-Q water after each run. After rinsing,
we place the microchannel in a closed glass container under Milli-Q water for further use. To
ensure the results presented in this study are free from errors, we conduct three independent runs
for each of the cases considered in this analysis. The average uncertainties in the calculation of
filling length and velocity reported here are found to be approximately 10% and 12%, respectively.

3. Filling kinematics at interfacial scale


We make an attempt here to scrutinize the filling kinematics of immiscible binary fluids
of unmatched densities and viscosities into narrow fluidic confinements using an analytical
methodology, consistent with the continuum approach [35]. To be precise, we attempt to peruse
the accuracy of our analytical solutions with the experimental results via a comparison of
the filling time into the capillary. Following the approximation, as has been considered in the
literature, we ignore the effects associated with the discreteness due to fluid microstructures in our
theoretical analysis [36]. The excerpts of the analysis are given below for the sake of completeness.
In surface tension dominated regime,1 the viscous force has a role of controlling the filling
5
dynamics. Considering this aspect, we undertake an effort to discuss the filling kinematics in the
following way. The analysis presented here is restricted to flow through a parallel plate channel

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


..........................................................
having a comparable width and height (cf. figure 1c) in the paradigm of continuum theory. The
velocity of the power-law fluid in a rectangular cross-section parallel plate channel, as considered
in this study, takes the following form:
  
 y (n+1)/n   z (n+1)/n
u = umax 1− 1− , (3.1)
h b

where umax is the maximum velocity and given as umax = −K|dp/dx|1/n , and h, b
are the half-height and half-width of the channel, respectively. While the parameter K
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

([kg]−1/n [m](2/n)+1 [s](2/n)−1 ) is a function of m and n, the magnitude of this parameter, which
is calculated numerically using a double integration method, depends on the height (2 h) and
width (2b) of the channel as well. For n = 1, i.e. for Newtonian fluids, and the channel dimensions
considered in this analysis, the value of K becomes 1.97 × 10−5 kg−1 m3 s. It may be mentioned
here that the calculated value is closer to the analytical value of 2 × 10−5 kg−1 m3 s for the
given dimensions of the channel. For the channel dimensions and the chosen values of CMC
concentration, 0.2% (n = 0.9, m = 0.01) and 0.4% (n = 0.82, m = 0.025), the values of K become
1.0024 × 10−6 and 2.0031 × 10−7 , respectively. Note that units of K and m are already mentioned
before. Also, it is worth mentioning that the approximate velocity profile delineated in equation
(3.1) reproduces the reported expression of the velocity profile for a case when b  h [30]. Now,
the flow rate of the power-law fluid through the parallel plate capillary can be written as




dp
1/n (n + 1)2
Q = 4hb K

. (3.2)
dx (2n + 1)2

Since we have considered the filling of incompressible power-law fluids, the flow rate Q is
constant. We integrate equation (3.2) primarily to obtain the pressure drop across the length of
the advancing fluid column (equivalently, length of the filled portion of the channel, i.e. at the
time ‘t’, the interface moves a distance x) as
  n
(2n + 1)2 Q
p = x. (3.3)
4hb(n + 1)2 K

The height of the microchannel restricts the effect of gravity on the underlying filling
phenomenon, and hence the pressure drop (left-hand side of equation (3.3)) corresponds to
the Laplace pressure p = 2σ (b + h)cosθ s /bh. Here, σ is the liquid–air surface tension and θ s
is the contact angle formed between the meniscus (liquid–air interface) and the channel wall.
It is essential to mention that in the surface tension modulated regime (effect of gravity is
insignificant), our analysis of filling kinematics is controlled by the balance between the capillary
and viscous pressure gradients [37,38]. Accordingly, the axial velocity of the meniscus can be
written as: dx/dt = Q/(2h × 2b). Now, equation (3.2) can be re-written in the following form:
n n
dx K α(1 + 1/n)2n 1
= , (3.4)
dt 4 (1 + 1/2n)2n x

where α = 2σ (b + h)cosθ s /bh. It is to be noted that equation (3.4) governs the filling characteristics
of power-law fluid in the microchannel, where the effect of gravity is negligibly small. For the

1
Surface tension force, which dominates over other forces driving imbibition such as gravity, is compared with the viscous
force to assess the filling kinematics in the capillary.
0.8
present experiment 6
Digilov [9]

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


..........................................................
0.6

0.4
l¯f

0.2
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

0
0 0.2 0.4 0.6

Figure 2. Plot of the variation of the temporal advancement of filling length into the capillary. A comparison between the
present experimental results with the results obtained from the Washburn analysis of Digilov [9] is presented. The markers are
used to represent the current experimental results, while the dotted line indicates the reported results [9]. (Online version in
colour.)

particular case of the uniform section, upon integrating equation (3.4), we get the following:

K (1 + 1/n)3 1/n
x(1+1/n) ∝ α t. (3.5)
n (2 + 1/n)2

It may be mentioned that equation (3.5) describes the filling kinematics in situations prevailing
with the compatible effect of surface tension and viscous forces. Having analysed equation
(3.5), one can find that for the Newtonian fluid (n = 1), the filling length follows the relation
x2 = (8K/9)αt, which is analogous to the classical Lucas–Washburn equation for a Newtonian
fluid. Having known the initial location of the interface, we solve equation (3.5) to obtain the
temporal advancement of the fluid in the capillary (filling into the capillary). As mentioned before,
we have used microchannels fabricated on a PMMA substrate for the experiments. It may be
mentioned here that the static contact angle of DI water with untreated PMMA is θs ∼ 67◦ , while
for an aqueous solution of CMC, θs ∼ 70◦ as reported in the literature [39,40]. Equation (3.5), which
conveys information about the filling kinematics, is solved to determine both the qualitative as
well as the quantitative spatio-temporal variation of the interface as demonstrated graphically in
the subsequent sections of this paper.
Before moving to the next section, we substantiate the applicability of the Washburn analysis,
as demonstrated above in predicting the experimental results. Washburn was the first to analyse
the imbibition phenomenon for Newtonian fluids in a capillary [1]. Following the reported
analysis [1], fluid is imbibed in a capillary upon experiencing a balance between the gradient of
capillary pressure and viscous stresses. At the same time, the effect of gravity is insignificant. It is
worth adding here that, as reported in the literature [9], the analysis also works well in describing
the filling of non-Newtonian fluids in a vertical capillary during the early time of filling since
the effect of gravity becomes trivial in this filling regime. Notably, in the reported analysis of
Digilov [9], the theoretical predictions from the Washburn analysis have been well validated by
the experimental results. Through the present analysis, by comparing our experimental results
with the results of Digilov [9] in figure 2, we re-establish the applicability of the Washburn analysis
in predicting the filling of non-Newtonian fluids in a horizontal capillary, wherein the effect of
gravity is insignificant.
Plots in figure 2 show the temporal variation of the filling length (dimensionless) l̄f in the
7
early regime of filling. The markers represent the current experimental results, while the dotted
line indicates the reported results [9]. It may be mentioned here that in the reported analysis of

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


..........................................................
Digilov [9], a vertical capillary was considered. To obtain this comparison, we consider aqueous
CMC solutions with 1% CMC. During the early instants of time, the temporal advancement of the
filling length in the horizontal capillary, as obtained from the current experimental investigations,
shows closer similarity with that of the vertical capillary. This inference signifies the insignificant
effect of gravity on the capillary filling kinematics and re-confirms the potential applicability of
the Washburn analysis in predicting the imbibition phenomenon of non-Newtonian fluid in a
horizontal capillary as well.

4. Results and discussion


Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

In line with the aim of the present study, we first discuss in §4a, the surface tension dominated
(surface tension force dominates over the other forces like gravity) filling characteristics, obtained
from the classical Washburn analysis [1,9], and compare those with our experimental results for
the pertinent cases. In §4b, we demonstrate experimental results for different cases considered
in this analysis. As the flow characteristics depend intricately on the participating forces
(capillary force and viscous force) in the spontaneous imbibition regime, the variations in the
magnitude of these forces affect the flow in the domain essentially to retain the balance between
capillary pressure gradient and viscous stresses. We discuss this aspect in greater detail from the
perspective of contact line dynamics in §4b. Finally, in §4c, we discuss the scaling analysis using
the experimental data.

(a) Comparison between theoretical and experimental results


We first discuss the cases where the filling characteristics predicted by the theoretical analysis
are compared with our experimental results. Accordingly, we show in figure 3a the temporal
variation of the length of fluid in the capillary, obtained for two different aqueous CMC
solutions, containing 0.2% cellulose (flow behaviour index n = 0.9, flow consistency index
m = 0.01 Pa sn ) and 0.4% cellulose (flow behaviour index n = 0.82, flow consistency index
m = 0.025 Pa sn ), respectively. We compare the theoretical results obtained from equation (3.5)
vis-à-vis experimental results for both the cases considered.
Having a closer look at figure 3a, one can find that the filling characteristics predicted
by the fluid kinematics at the interfacial scale qualitatively endorse the results obtained from
experiments for both the cases considered. It is noteworthy to mention that the trends of the
temporal variation of the capillary filling recorded from our experiments are captured by the
theoretical calculations in an accurate manner. Next, coming to the quantitative evaluation of
the temporal variation of filling length of the advancing fluid (fluid under focus here) in the
capillary with a change in CMC concentration. We observe from figure 3a that an increase in
cellulose concentration in aqueous solution (flow behaviour index decreases and fluid becomes
more shear-thinning in nature) slows down the filling rate.
We further make an effort to explain this rheology modulated variation in filling characteristics
by looking at the fluid kinematics at interfacial scale under the modulation of surface
tension effect. Albeit, this aspect has been explained earlier in §3, we briefly reiterate the
same as discussed next. We mention here that the shear rate at the walls of the channel
is an important parameter for this analysis in order to figure out the physical reasoning
of the filling kinematics. Now using the expression of the flow rate through the channel
Q = 4(hb)dx/dt and equation (3.4), we can write the shear rate (average) at the walls as γ̇wall, avg =
2K((b/h + h/b)/(b + h))((n + 1)2 /n(2n + 1))(α/x)1/n . From this expression, it is apparent that the
shear rate will decrease with increasing x regardless of the value of n, but the form of
decrease will depend on n as can be seen from the expression of apparent viscosity given
next. To be noted is that the apparent viscosity of the non-Newtonian fluid following the
(a) 1800 (b) 1800
height = 400 µm 8
experimental height = 400 µm
experimental theoretical: Carreau model
theoretical: power-law model

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


1400 1400

..........................................................
n = 0.9

1000 1000
lf (µm)

lf (µm)
water-CMC 0.2%
(m = 0.010; n = 0.9)
water-CMC 0.4%
600 (m = 0.025; n = 0.82) 600 n = 0.82

200 200

0 0
20 40 60 20 40 60
tf (s)
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

tf (s)

Figure 3. (a) Plot showing the capillary filling length (lf ) versus time (tf ) for two different inelastic non-Newtonian fluids. Two
different non-Newtonian (shear-thinning) fluids have been considered for estimating the filling dynamics in the capillary which
are aqueous carboxy-methyl-cellulose solutions, containing 0.2% cellulose (flow behaviour index n = 0.9, flow consistency
index m = 0.01) and 0.4% cellulose (flow behaviour index n = 0.82, flow consistency index m = 0.025 Pa sn ), respectively.
Analytical results obtained from the filling kinematics in the capillary are compared with the experimental results for both
cases, and (b) shows the variation of the temporal advancement of filling length (lf ) into the capillary. A comparison between
the present experimental results with the theoretical predictions by the Carreau–Yasuda model [41] is presented. The markers
are used to represent the current experimental results, while the dotted line indicates the theoretical results, obtained using the
Carreau–Yasuda model [41]. (Online version in colour.)

expression of the shear rate at the walls can be written as μapp = mK1 (α/x)(1−1/n) , where
K1 = 2K((b/h + h/b)/(b + h)){(n + 1)2 /n(2n + 1)2 }. As evident from this expression, the spatial as
well as the temporal variation (using equation (3.5)) of apparent viscosity are governed by
the following relations: x−(1−1/n) and t(1−n)/(1+n) , respectively. The above relationship is also
suggestive of the following: as the inelastic non-Newtonian fluid imbibes into the channel, the
apparent viscosity increases with time for the class of shear-thinning fluids (n < 1). Also, the
increment in apparent viscosity of the inelastic non-Newtonian fluid is higher for the fluids
having higher shear-thinning nature, i.e. for the lesser value of n as it progresses through the
capillary. It may be mentioned here that the fluid with more shear-thinning nature is more viscous
as well since m increases (cf. table 1). With increasing m, the apparent viscosity becomes higher,
as can be verified from the definition of apparent viscosity given above. Quite notable, for n = 1
(Newtonian fluid), viscosity assumes a constant value, and the temporal advancement of the fluid
in the capillary follows the well-established Lucas–Washburn dynamics x ∼ t1/2 [1,9,42].
The variation of apparent viscosity with flow behaviour index (n), as elaborated above, follows
from the variation of wall shear rate (γ̇wall ) with x and works well for the pseudo-steady-state
movement of the interface in the capillary. Quite notably, a pseudo-steady-state movement of the
interface (formed between aqueous CMC solution and air) in the capillary is verified through
the variations depicted in figure 3a. Taking a note on this aspect of increasing apparent viscosity
with increasing shear-thinning nature (realized by a smaller value of n as well as higher value of
m in table 1) of fluids being imbibed into the capillary, it can be argued that a relatively larger
resistance offered to the interface of aqueous CMC solution containing 0.4% CMC for a given
strength of capillary pressure gradient slows down the filling rate. This observation is mostly
reflected in figure 3a. Also, we mention another relevant point in this context. At the beginning of
the filling, predicted values by the theoretical analysis differ significantly from the experimental
observations. We attribute this observation to the inherent drawback of the power-law model,
used here to describe the rheology of the inelastic non-Newtonian fluid. It is important to mention
that, for a very high shear rate developed at the earlier instants of filling, the power-law model
yields a complication of the infinite apparent viscosity prediction. In particular, the power-law
model cannot capture a very high shear rate developed at the earlier instants of filling. Due to
9
this limitation of the power-law model, we observe a notable difference between the experimental
results and theoretical prediction of the filling length at initial temporal instants, as evidently

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


..........................................................
verified in figure 3a.
To check the correctness of our experimental results as well as to establish the argument of
the infinite viscosity problem of the power-law model at a high shear rate, we make an effort
in figure 3b to compare the experimental results with the theoretical predictions obtained by
extending the analysis to the Carreau–Yasuda model [41]. Note that the Carreau–Yasuda model
does not have the zero-shear rate complication. We use the rheological data of CMC solutions to
obtain model parameters of the Carreau–Yasuda model, and subsequently to plot the variations
depicted in figure 3b. Note that the value of the other model parameters conforms to the reported
value in the referred literature [41]. Pertinent analysis of the underlying filling kinematics of the
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

Carreau–Yasuda model [41] in a parallel plate channel (b  h) is outlined in appendix A.

(b) Effect of variation in cellulose composition on filling: contact line dynamics


The temporal variation of filling length with a change in CMC composition is essentially a
reflection of the effect of apparent viscosity on the capillary filling dynamics, as discussed in the
preceding section. To have a better understanding of this aspect, we here look at this phenomenon
in greater detail from the perspective of the variation of contact line motion. The relatively
higher apparent viscosity with increasing shear-thinning nature of the fluid slows down the
filling process as elaborated earlier. Here, in the inset of figure 4a, we also show the experimental
observations on filling length of a Newtonian fluid (DI water), keeping the common parameters
constant. On account of a relatively lesser viscous resistance, the advancing front of the DI water
penetrates a larger distance at a given time into the capillary, as evident from the variation shown
in the inset. Next, we look at the variation of the contact line velocity portrayed in figure 4b
for two different cases, as mentioned in the figure. It is observed in figure 4b that the contact
line velocity (vmcl ) is less for the fluid having higher shear-thinning nature (n = 0.82). As already
mentioned, in the present endeavour, we make an effort to explore the effect of fluid rheology;
to be precise, the shear-thinning effect on the underlying filling characteristics in a microfluidic
channel under the influence of surface tension force. Since the advancing liquid is imbibed into
the capillary following the balance between the surface tension force and viscous resistance, the
three-phase contact line advances faster than the movement of the interface at the middle of
the channel. As evident from the snapshots presented in figure 4c,d, the advancing fluid, front
following the imbibition dynamics takes a convex shape. Note that the slower movement of the
contact line (higher viscous resistance due to relatively higher apparent viscosity with lesser n and
higher m) for the case when a relatively higher shear-thinning fluid displaces air into the capillary
(cf. figure 4b) slows down the movement of the meniscus along the length of the channel. This
phenomenon results in a reduction in filling rate, as confirmed in figure 4a.
We further focus on the variation of contact line motion, which plays a vital role in the capillary
imbibition dynamics in the present scenario, from the perspective of the energy budget of the
moving contact line. It may be mentioned here that as the contact line moves along the channel, it
encounters a competitive environment from the forces acting on it. It is important to mention that
participating forces are stemming from two different sources of energy: dissipation due to bulk
viscosity Dvis and kinetic energy Dkin [43,44]. The expression of the shear dissipation due to bulk

viscosity is Dvis = ∀ μapp (∇u + ∇uT ) : (∇u + ∇uT )dV and the rate of change of the total kinetic

energy becomes Dkin = ∀ 0.5ρ(∂u2 /∂t) dV. As can be seen, the dissipation due to bulk viscosity
scales with apparent viscosity of the fluid μapp , while the rate of change of the kinetic energy
scales with the fluid density. The kinetic energy provides the required energy for the movement
of the contact line in the process, while the dissipation due to apparent viscosity resists the contact
line motion [43].
Following this effect, a relatively higher magnitude of Dvis as expected for a fluid having
higher apparent viscosity, precisely for higher shear-thinning nature (lesser n and higher m),
(a) (b)
1800
10
40
25 000
DI water
height = 400 mm
20 000

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


lf (µm)

..........................................................
15 000
1400 10 000
30
5000
water-CMC 0.4%

umcl (µm s–1)


0 20 40 60
t f (s) (m = 0.025; n = 0.82)
lf (µm)

1000 water-CMC 0.2%


water-CMC 0.2% 20 (m = 0.010; n = 0.9)
(m = 0.010; n = 0.9)
water-CMC 0.4%
600 (m = 0.025; n = 0.82)

10

200

0 20 40 60 0 20 40 60
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

tf (s) t (s)

(c) (d)
water-CMC air 500 mm
interface

0.2% carboxy-methyl-cellulose (CMC) 0.4% carboxy-methyl-cellulose (CMC)

Figure 4. (a) Plot showing the capillary filling length (lf ) versus time (tf ) for two different cases of carboxy-methyl-cellulose
concentration (inset shows the temporal variation of filling length for Newtonian fluid). (b) Variation of contact line velocity
as the aqueous solutions with 0.2% cellulose (flow behaviour index n = 0.9, flow consistency index m = 0.01 Pa sn ) and
0.4% cellulose (flow behaviour index n = 0.82, flow consistency index m = 0.025 Pa sn ), respectively, advance in the capillary.
(c) Snapshot of the interface (formed between aqueous carboxy-methyl-cellulose solution and air) captured at a temporal
instant t = 20s for 0.2% cellulose in solution and (d) snapshot of the interface (formed between aqueous carboxy-methyl-
cellulose solution and air) captured at a temporal instant t = 40 s for 0.2% cellulose in solution. (Online version in colour.)

reduces the contact line velocity, as confirmed in figure 4b. For the purpose of completeness, we
may mention that the contact line velocity is significantly less for advancing fluid with 0.4% CMC
(flow behaviour index n = 0.82, flow consistency index m = 0.025 Pa sn ) than that for the aqueous
solution with 0.2% CMC (cf. figure 4b). The lesser contact line velocity eventually delays the
filling dynamics for aqueous solutions with 0.4% cellulose in the surface tension driven regime, as
witnessed by figure 4 and images presented therein. It is worth mentioning that the morphological
irregularities at the channel wall, which are inevitable with the microfabrication processes, lead
to the appearance of oscillations on the velocity variations delineated in figure 4b.

(c) Scaling analysis


Apart from the temporal variations of the filling length into the capillary depicted above, we
make an effort to estimate the filling regime through scaling analysis. Important to mention,
we focus our attention on the late stage of the filling process, since the predicted behaviour of
the filling dynamics at the early stages, owing to the model considered in this analysis, is not
adequate. In an effort to establish the scale which governs the filling dynamics at later stages,
we compare the surface tension forcing to the viscous force. As seen from figure 5, the late-stage

filling closely follows x ∼ t, and we call it the modified Washburn regime2 [1]. We discuss this
scaling argument as elaborated next. While performing the scaling estimation for the late-stage
regime, we consider both the cases of the aqueous CMC solution having 0.2% and 0.4% CMC
concentration. By comparing the viscous force fvis ∼ m(xn+1 2(b + h)/tn hn ) with the surface tension

2
We call it the modified Washburn regime as the Washburn equation is used to calculate filling of non-Newtonian fluid.
water-CMC 0.4% (m = 0.025; n = 0.82) 11
103

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


..........................................................
x (µm) water-CMC 0.4% (m = 0.025; n = 0.82)
102

experimental
scaling: x ~ ÷ t
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

10 102
t (s)

Figure 5. Plot of scaling analysis of the results obtained from experimental investigations during late-stage filling. The markers
represent the experimental results, whereas the dotted line represents the results obtained from the scaling analysis. The later
stage of the filling process of non-Newtonian fluid, where surface tension force and viscous force are the participating forces for
the interface movement, follows modified Washburn dynamics [1]. (Online version in colour.)

force fsur ∼ 4σ b(1 + h/b), we get: x ∼ (2σ hn /m)1/(n+1) tn/(n+1) since uavg ∼ x/t. For the Newtonian
 √
fluids (n = 1), this scale x ∼ t confirms the well-known Lucas–Washburn regime.

5. Conclusion
We have experimentally investigated capillary filling characteristics of inelastic non-Newtonian
fluids in the regime compatible with a zone with balanced capillary and viscous forces. To
represent the rheology of the inelastic non-Newtonian fluids, we have used the Ostwald–de
Waele power-law model. Aqueous solutions of CMC with different concentrations (0.2% and
0.4% CMC by weight), which exhibit shear-thinning behaviour, are used as the working fluids
in this analysis. We have also established a functional relationship between the advancement
of the filling fluid into the capillary and the corresponding time through a theoretical analysis
that is consistent with the fluid kinematics at the interfacial scale. Theoretical prediction of the
filling length into the capillary corroborates our experimental results obtained for two different
cellulose (CMC) concentrations in aqueous solution. Depicting the temporal variation of the
filling length for different parameters, we have unveiled that the dynamics of contact line motion,
conforming to the mass conservation principle, plays an important role in the movement of the
advancing fluid meniscus in the channel. We have demonstrated that the motion of the interface
brings about control over the filling length of the advancing fluid. Quite importantly, through
our theoretical analysis, as evidenced by the experimental insights, we have established that the
power-law model cannot capture a very high shear rate (infinite effective viscosity prediction)
developed at the earlier instants of filling. By contrast, calculating the filling length from the fluid
kinematics, and subsequent comparison with the experimental data, we have been able to show
that the Carreau–Yasuda model can perfectly reproduce the phenomenon even during high shear
rate condition. Also, by comparing our experimental results with the results of Digilov [9], we
have re-established that in the regime of insignificant gravitational effect, the Washburn equation
perfectly works to predict the filling characteristics of non-Newtonian fluids. In addition to the
theoretical predictions, we have shown through a scaling analysis that the late-stage filling of

shear-thinning fluids closely follows the variation of x ∼ t, and we refer to this regime as the
extended Washburn regime. We believe that the inferences of the present analysis are likely to
12
provide useful input for designing portable on-chip devices typically used for clinical diagnostics.

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


Data accessibility. Available as the electronic supplementary material, Data.

..........................................................
Authors’ contributions. The conceptualization of the problem including the theoretical foundation was conceived
by P.K.M. and G.B. The experiments were performed by S.R.G. and S.K.M. Theoretical calculations, figures
and graphics were prepared by all the authors. The outline of the manuscript along with its organization was
structured by P.K.M. The interpretation of the results and the preparation of the manuscript were done by
P.K.M. and G.B. The research was supervised and directed by P.K.M. and G.B. All authors gave final approval
for publication and agree to be held accountable for the work performed therein.
Competing interests. We declare we have no competing interests.
Funding. No funding has been received for this article.
Acknowledgements. Authors gratefully acknowledge the financial support provided by the SERB (DST), India,
through project no. ECR/2016/000702/ES. Authors would like to acknowledge Mr Sudip Shyam and Mr
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

Harshad Gaikwad, senior research scholars of Mechanical Engineering Department, Indian Institute of
Technology Guwahati, for their help during experiments pertaining to the data analysis. One of us (G.B.)
acknowledges his gratitude for the J.C. Bose National Fellowship of SERB (DST), India.

Appendix A. Filling kinematics using the Carreau–Yasuda model


The flow velocity in a parallel plate channel (b > h) following the Carreau model [45] takes the
following form:   y (n+1)/n 
u = umax 1 − . (A 1)
h
The expression of umax is given below:
 1/n
n 1
umax = |dp/dx|1/n . (A 2)
(n + 1) {(μ0 /h2 )(λ/h)n−1 }

The following equation describes the velocity of the meniscus:


n  n+1 
dx h α 1
= . (A 3)
dt bn {2n μo λn−1 } x

The equation can describe the filling kinematics under the surface tension effect as
   
1/n h(n+1)/n α 1/n
x dx = 1/n
dt. (A 4)
2b {μ0 λ(n−1)/n }

References
1. Washburn EW. 1921 The dynamics of capillary flow. Phys. Rev. 17, 273–283. (doi:10.1103/
PhysRev.17.273)
2. Kusumaatmaja H, Pooley CM, Girardo S, Pisignano D, Yeomans JM. 2008 Capillary filling in
patterned channels. Phys. Rev. E 77, 067301. (doi:10.1103/PhysRevE.77.067301)
3. Huang W, Liu Q, Li Y. 2006 Capillary filling flows inside patterned-surface microchannels.
Chem. Eng. Technol. 29, 716–723. (doi:10.1002/ceat.200500332)
4. Mognetti BM, Yeomans JM. 2009 Capillary filling in microchannels patterned by posts. Phys.
Rev. E 80, 056309. (doi:10.1103/PhysRevE.80.056309)
5. Gaikwad HS, Roy A, Mondal PK. 2020 Autonomous filling of a viscoelastic fluid in a
microfluidic channel: effect of streaming potential. J. Non–Newton. Fluid Mech. 282, 104317.
(doi:10.1016/j.jnnfm.2020.104317)
6. Das S, Mitra SK. 2013 Different regimes in vertical capillary filling. Phys. Rev. E 87, 063005.
(doi:10.1103/PhysRevE.87.063005)
7. Das S, Waghmare PR, Mitra SK. 2012 Early regimes of capillary filling. Phys. Rev. E 86, 067301.
(doi:10.1103/PhysRevE.86.067301)
8. Shardt O, Waghmare PR, Derksen JJ, Mitra SK. 2014 Inertial rise in short capillary tubes. RSC
Adv. 4, 14 781–14 785. (doi:10.1039/c4ra00580e)
9. Digilov RM. 2008 Capillary rise of a non-Newtonian power law liquid: impact of the fluid
13
rheology and dynamic contact angle. Langmuir 24, 13 663–13 667. (doi:10.1021/la801807j)
10. Lin Y, Zheng L, Zhang X. 2013 Magnetohydrodynamics thermocapillary Marangoni

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


..........................................................
convection heat transfer of power-law fluids driven by temperature gradient. J. Heat Transfer
135, 051702. (doi:10.1115/1.4023394)
11. Mondal PK, Ghosh U, Bandopadhyay A, DasGupta D, Chakraborty S. 2013 Electric-field-
driven contact-line dynamics of two immiscible fluids over chemically patterned surfaces in
narrow confinements. Phys. Rev. E 88, 023022. (doi:10.1103/PhysRevE.88.023022)
12. Gorthi SR, Mondal PK, Biswas G. 2017 Magnetic-field-driven alteration in capillary filling
dynamics in a narrow fluidic channel. Phys. Rev. E 96, 013113. (doi:10.1103/PhysRevE.
96.013113)
13. Gorthi SR, Gaikwad HS, Mondal PK, Biswas G. 2020 Surface tension driven filling in
a soft microchannel: role of streaming potential. Ind. Eng. Chem. Res. 59, 3839–3853.
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

(doi:10.1021/acs.iecr.9b00767)
14. Jacqmin D. 2000 Contact-line dynamics of a diffuse fluid interface. J. Fluid Mech. 402, 57–88.
(doi:10.1017/S0022112099006874)
15. DasGupta D, Mondal PK, Chakraborty S. 2014 Thermocapillary-actuated contact-line motion
of immiscible binary fluids over substrates with patterned wettability in narrow confinement.
Phys. Rev. E 90, 023011. (doi:10.1103/PhysRevE.90.023011)
16. Mondal PK, DasGupta D, Bandopadhyay A, Chakraborty S. 2014 Pulsating flow driven
alteration in moving contact-line dynamics on surfaces with patterned wettability gradients.
J. Appl. Phys. 116, 084302. (doi:10.1063/1.4893705)
17. Sheng P, Zhou M. 1992 Immiscible-fluid displacement: contact-line dynamics and the
velocity-dependent capillary pressure. Phys. Rev. A 45, 5694–5708. (doi:10.1103/PhysRevA.
45.5694)
18. Tung C, Krupa O, Apaydin E, Liou J-J, Diaz-Santana A, Kim BJ, Wu M. 2013 A contact line
pinning based microfluidic platform for modelling physiological flows. Lab Chip 13, 3876–
3885. (doi:10.1039/c3lc50489a)
19. Stone HAA, Stroock ADD, Ajdari A. 2004 Engineering flows in small devices. Annu. Rev. Fluid
Mech. 36, 381–411. (doi:10.1146/annurev.fluid.36.050802.122124)
20. Kim HJ, Jang WK, Kim BH, Seo YH. 2016 Advancing liquid front shape control in capillary
filling of microchannel via arrangement of microposts for microfluidic biomedical sensors.
Int. J. Precis. Eng. Manuf. 17, 59–63. (doi:10.1007/s12541-016-0008-x)
21. Kusumaatmaja H, Yeomans JM. 2007 Modeling contact angle hysteresis on chemically
patterned and superhydrophobic surfaces. Langmuir 23, 6019–6032. (doi:10.1021/la063218t)
22. Bandopadhyay A, Mandal S, Chakraborty S. 2016 Streaming potential-modulated capillary
filling dynamics of immiscible fluids. Soft Matter 12, 2056–2065. (doi:10.1039/C5SM02687C)
23. Zhou J, Khodakov DA, Ellis AV, Voelcker NH. 2012 Surface modification for PDMS-based
microfluidic devices. Electrophoresis 33, 89–104. (doi:10.1002/elps.201100482)
24. Li X, Tian J, Nguyen T, Shen W. 2008 Paper-based microfluidicdevices by plasma treatment.
Anal. Chem. 80, 9131–9134. (doi:10.1021/ac801729t)
25. Mondal PK, Ghosh U, Bandopadhyay A, DasGupta D, Chakraborty S. 2014 Pulsating
electric field modulated contact line dynamics of immiscible binary systems in narrow
confinements under an electrical double layer phenomenon. Soft Matter 42, 8512–8523.
(doi:10.1039/C4SM01583E)
26. Darhuber AA, Valentino JP, Davis JM, Troian SM, Wagner S. 2003 Microfluidic actuation by
modulation of surface stresses. Appl. Phys. Lett. 82, 657–659. (doi:10.1063/1.1537512)
27. Gorthi SR, Mondal PK, Biswas G, Sahu KC. In press. Electro-capillary filling in a microchannel
under the influence of magnetic and electric fields. Can. J. Chem. Eng. (doi:10.1002/cjce.23876)
28. Som SK. 1983 Theoretical and experimental studies on the coefficient of discharge and spray
cone angle of a swirl spray pressure nozzle using a power-law non-Newtonian fluid. J. Non-
Newton. Fluid Mech. 12, 39–68. (doi:10.1016/0377-0257(83)80004-4)
29. Mondal PK, Dasgupta D, Chakraborty S. 2015 Rheology-modulated contact line dynamics
of an immiscible binary system under electrical double layer phenomena. Soft Matter 11,
6692–6702. (doi:10.1039/c5sm01175b)
30. Sarma R, Gaikwad H, Mondal PK. 2017 Effect of conjugate heat transfer on entropy generation
in slip-driven microflow of power law fluids. Nanoscale Microscale Thermophys. Eng. 21, 26–44.
(doi:10.1080/15567265.2016.1272655)
31. de Waele A. 1923 Viscometry and plastometry. J Oil Colour Chem. Assoc. 6, 33–69.
14
32. Ostwald W. 1925 Ueber die Geschwindigkeitsfunktion der Viskosität disperser Systeme. I.
Kolloid-Zeitschrift 36, 99–117. (doi:10.1007/BF01431449)

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 476: 20200496


..........................................................
33. Metzner AB. 1953 Advances in chemical engineering. New York, NY: Academic Press.
34. Som SK, Biswas G. 1984 Initiation of air core in a swirl nozzle using time-independent power-
law fluids. Acta Mech. 51, 179–197. (doi:10.1007/BF01177071)
35. Fries N, Dreyer M. 2008 The transition from inertial to viscous flow in capillary rise. J. Colloid
Interface Sci. 327, 125–128. (doi:10.1016/j.jcis.2008.08.018)
36. Holloway W, Aristoff JM, Stone HA. 2011 Imbibition of concentrated suspensions in
capillaries. Phys. Fluids 23, 081701. (doi:10.1063/1.3619217)
37. Xu B, Ooi KT, Mavriplis C, Zaghloul ME. 2003 Evaluation of viscous dissipation in liquid flow
in microchannels. J. Micromech. Microeng. 13, 53–57. (doi:10.1088/0960-1317/13/1/308)
38. Koo J, Kleinstreuer C. 2004 Viscous dissipation effects in microtubes and microchannels. Int.
Downloaded from https://royalsocietypublishing.org/ on 07 July 2023

J. Heat Mass Transf. 47, 3159–3169. (doi:10.1016/j.ijheatmasstransfer.2004.02.017)


39. Ma Y, Cao X, Feng X, Ma Y, Zou H. 2007 Fabrication of super-hydrophobic film from
PMMA with intrinsic water contact angle below 90°. Polymer 48, 7455–7460. (doi:10.1016/j.
polymer.2007.10.038)
40. Tokuda K, Ogino T, Kotera M, Nishino T. 2015 Simple method for lowering poly(methyl
methacrylate) surface energy with fluorination. Polym. J. 47, 66–70. (doi:10.1038/pj.2014.91)
41. Cho YI, Kensey KR. 1991 Effects of the non-Newtonian viscosity of blood on flows in
a diseased arterial vessel. Part 1: steady flows. Biorheology 28, 241–262. (doi:10.3233/BIR-
1991-283-415)
42. Fisher LR, Lark PD. 1979 An experimental study of the Washburn equation for liquid flow in
very fine capillaries. J. Colloid Interface Sci. 69, 486–492. (doi:10.1016/0021-9797(79)90138-3)
43. Mondal PK, Chaudhry S. 2018 Effects of gravity on the thermo-hydrodynamics of moving
contact lines. Phys. Fluids 30, 042109. (doi:10.1063/1.5017937)
44. Wang X-P, Qian T, Sheng P. 2008 Moving contact line on chemically patterned surfaces.
J. Fluid Mech. 605, 59–78. (doi:10.1017/S0022112008001456)
45. Zhao C, Yang C. 2011 Electro-osmotic mobility of non-Newtonian fluids. Biomicrofluidics 5,
14110. (doi:10.1063/1.3571278)

You might also like