Vous êtes sur la page 1sur 9

Microfluid Nanofluid (2016) 20:63

DOI 10.1007/s10404-016-1726-1

RESEARCH PAPER

Controlling capillarydriven surface flow ona paperbased


microfluidic channel
JoelSongok1 MarttiToivakka1

Received: 21 November 2015 / Accepted: 3 March 2016 / Published online: 30 March 2016
Springer-Verlag Berlin Heidelberg 2016

Abstract This paper describes two methods for controlling capillary-driven liquid flow on microfluidic channels.
Unlike flow driven by external forces, capillary-driven flow
is dominated by interfacial phenomena and, therefore, is
sensitive to the channel geometry and chemical composition (surface energy) along the channel. The first method
to control fluid flow is based on altering surface energy
along the channel through regulation of UV irradiation
time, which enables adjusting the contact angle along the
fluid path. The slowing down (delay) of the liquid flow
depends on the stripe length and its position in the channel.
Using this technique, we generated flow delays spanning
from a second to over 3min. In the second approach, we
manipulated the flow velocity by introducing contractions
and expansions in the channel. The methods used herein
are inexpensive and can be incorporated to the microfluidic
channel fabrication step. They are capable of controlling
liquid flow with precise time delays without introducing
the foreign matter in the fluidic device.
Keywords Capillary forces Paper-based microfluidics
Surface flow Control

Electronic supplementary material The online version of this


article (doi:10.1007/s10404-016-1726-1) contains supplementary
material, which is available to authorized users.
* Joel Songok
joel.songok@abo.fi
1

Laboratory ofPaper Coating andConverting andCenter


forFunctional Materials, Abo Akademi University,
Porthaninkatu 3, 20500bo/Turku, Finland

1Introduction
Recently, paper has attracted interest as a possible substitute for conventional microfluidics substrates because it is
cost-effective, user-friendly, environmentally friendly, and
simple to produce and dispose off. Paper-based microfluidic devices (PADs) enable precise transport of microliter volumes of liquid without pumps driven by capillary
action. Accordingly, it has great potential for point-of-care
(POC) diagnostics in resource-limited and home care settings (Hitzbleck etal. 2012; Von Lode 2005). PADs can
perform fluid handling and quantitative analysis, and
numerous applications have already been demonstrated in
healthcare, veterinary medicine, environmental monitoring,
and food safety: recent reviews detail specific applications
(Ballerini etal. 2012; Cate etal. 2015; He etal. 2015; Li
etal. 2012; Sackmann etal. 2014; Santhiago etal. 2014;
Yetisen etal. 2013). Controlled liquid transport in devices
is crucial for obtaining reproducible results and consistent
sensitivity (Yetisen etal. 2013). Active metering techniques
such as mechanical pumping, electro-wetting, magnetic
field, and electro-osmotic force provide rapid and precise
manipulation of microliter volumes of liquid. However,
they require additional energy sources or equipment (e.g.,
voltage amplifier, pump, power supplier) hence leading to
increased size, cost and complexity of microfluidic device.
The development of ways to control, and particularly delay
the fluid flow would enable greater functionalities to be
implemented on paper-based microfluidic devices. For
instance, the introduction of control over liquid transport
could enable sequential delivery of samples to perform
multiple timed analytical steps in a single step.
Several ways, including photolithography (Martinez
etal. 2008), laser techniques (Fu etal. 2010a, b), plasma
oxidation (Li etal. 2008), and different printing techniques,

13


63
Page 2 of 9

namely flexographic (Olkkonen etal. 2010), inkjet (Abe


etal. 2008, 2010), and wax printing (Carrilho etal. 2009;
Dungchai etal. 2011; Songjaroen etal. 2011; Zhong etal.
2012), have been used to fabricate channels for PADs.
In our earlier work (Songok etal. 2014), we reported a
new technique for creating channels that utilize the photocatalytic property of titanium dioxide coated on paper.
The technique presents two advantages when compared
to alternative methods. First, the flow of liquid occurs on
paper surface rather than in the paper matrix thus eliminating problems associated with flow in paper fiber network such as fiber swelling that introduces complexity in
the liquid flow, difficulty of transporting cells through the
paper matrix, reduction of paper assay effectiveness due to
the interaction of biological samples with cellulose fibers,
and mostly, liquid retention in the fiber matrix that leads
to large sample volumes (Ballerini etal. 2012; Cheng etal.
2010; Songjaroen etal. 2012; Thuo etal. 2014). Secondly,
the fabrication is carried out in two steps, of which the first
step, i.e., deposition of TiO2 on the paper surface is accomplished in a roll-to-roll process, which enables high-speed,
large-scale manufacturing.
Crucial to the operation of paper-based assays is the
ability to manipulate fluids for the precise timing of reagent delivery and controlling of reagent volumes. A number of studies have reported on different approaches for
manipulating delivery of fluids in paper-based devices. Li
etal. created a manually controlled switch that connects
and disconnects the channel (Li etal. 2008). The channel
geometry (i.e., the channel width and the length of the inlet
leg) were varied to accomplish different arrival time for
the liquid (Fu etal. 2010a, b, 2011; Osborn etal. 2010).
The Yager group has recently applied dissolvable sugar
(sucrose) onto the channel to create a tunable time delay
(Lutz etal. 2013). The delay time occurs as the liquid dissolves the sucrose along its path. The delay time depends
on the amount of dissolvable material deposited. Jahanshahi-Anbuhi etal. used water-soluble pullulan film that
automatically dissolves and eventually cuts off the flow to
the sensing zone (Jahanshahi-Anbuhi etal. 2014). Other
techniques involve modification of the wetting properties of
paper using hydrophobic material such as paraffin wax or
use of surfactant to design timed-opening valves or diodes
(Fu etal. 2010a, b; Lutz etal. 2013; Noh and Philips 2010a,
b) (Chen etal. 2012). Although these approaches successfully control the capillary flow rate, they may exhibit
some shortcomings. Methods that use manually controlled
switches require user intervention. The use of dissolvable
materials may interfere with the functionality of the device
and adds an additional processing step. Further, implementation of flow delay by changing the channel geometry,
e.g., by expanding the channel width or by increasing of
the length, increases the required reagent volumes.

13

Microfluid Nanofluid (2016) 20:63

In this work, we propose two simple ways to control


capillary-driven flow on paper surface. The first method
is based on altering the surface wettability of a section
across the flow path so as to produce a surface energy
barrier. This is a modification of the technique we have
earlier reported for creating microfluidics channels (Songok etal. 2014). The technique involves spraying TiO2
nanoparticles onto paperboard using liquid flame spray
treatment to form a superhydrophobic surface. A channel is created by exposing the TiO2-coated paper to UV
irradiation through a photomask, which creates a hydrophilic path for liquid confined by hydrophobic boundaries. We previously observed that by varying the exposure
time, surfaces with different surface energies (resulting
in different contact angles for liquids) can be created.
Therefore, to produce flow control, we took advantage
of surface energetics by introducing a patch of low surface energy (high contact angle) along a channel of high
surface energy through irradiating different sections of
the channel to UV for different times. Wetting behavior
of a liquid depends on the surface energy of the substrate
and surface tension of the liquid. At the liquidsubstrate
interface, if the adhesive forces are stronger than cohesive forces, then wetting of the substrate occurs and vice
versa. Here, the surface energy gradient is useful for
transporting small liquid volumes and to control fluid
flow. Stopping and initiating flow within the device eliminates the need for expensive pumps and controllers hence
potentially reducing the cost of current systems. The second approach utilizes contractions and expansions in the
channel to adjust the capillary spreading rate and delays.
The work described here aimed at introducing a new way
to create flow delays for liquid delivered on paper surface
without adding fabrication steps or introducing new material along the channel.

2Materials andmethods
Commercially available pigment-coated paperboard
(200g/m2, Stora Enso, Sweden) was coated with TiO2
nanoparticles in a roll-to-roll process using the liquid
flame spray (LFS) method. The LFS coating was carried out in the pilot scale paper converting line at Tampere University of Technology. A detailed description of
the process has been presented previously (Mkel etal.
2011; Teisala etal. 2010). The TiO2-coated paper was then
exposed to ultraviolet irradiation (320390nm) (Bluepoint
4 ecocure, Hnle UV technology, DE) via a photomask.
Figure 1 below shows the LFS coating process, a SEM
cross-sectional image of the coated paper and UV irradiation step. Fast and less-expensive photomask was made by
cutting office paper, and for high resolution a photomask

Page 3 of 9 63

Microfluid Nanofluid (2016) 20:63

O
H

Liquid
precursor

TiO nanoparticles
2

Pigment
coating
UV- irradiation

Photomask
Paper

2 mm
1 m

Base paper
2 m

Fig. 1a Liquid flame spray setup, b UV irradiation setup, c cross-sectional and top SEM images showing TiO2 LFS nanoparticles on paperboard and d the two types of photomasks used to block UV irradiation: silver ink printed on glass (top) and cut office paper (bottom)

3Results anddiscussion
3.1Paper surface treatment
Superhydrophobic paper surface was created by LFS coating of TiO2 nanoparticles. Figure2 shows the shift in the
water contact angle of paperboard from nearly 70 to 150
after TiO2 deposition. The rise in the contact angle is due
to a combination of the nanoparticle surface chemistry
and the complex surface structure that incorporates both
nano- and microscale roughness to enhance the water
repellency (Teisala etal. 2010, 2013). Exposure to UV
light transformed the surface wettability from superhydrophobic to hydrophilic regime. Depending on the exposure
time, the water contact angle of TiO2-coated paperboard
dropped from 150 to nearly 70, 30, and 10 after 5,
10, and 20min UV irradiation. This conversion has been

5 min
20 min

180

Contact angle [Deg]

made by printing silver ink on a glass slide using inkjet


printer (DimatixTM materials printer DMP-2831, Fujifilm,
USA) was applied. The various geometries and shapes are
presented in subsequent figures. The irradiation intensity
was kept constant at 50mWcm2, but the exposure times
were varied so as to create surfaces with different wettability. The surface wettability was characterized by contact
angle measurements (KSV CAM 200, KSV Instruments
Ltd., FI) using a Laplacian fit to the projected 6-l droplet
curvature.
The test sample was mounted horizontally on to a flat
surface using double-sided tape and a 20-l liquid droplet
was deposited on the sample point. The advancing meniscus position was recorded from the top view. The position
of the liquid front as a function of time was measured at
100 fps using a video camera. The video files were then
used to extract the distance travelled by the fluid front in
time. A small quantity of amaranth red dye was added to
the water sample to improve the visual contrast of the thin
liquid films. The results are presented as averages of 46
parallel measurements.
The system for demonstrating sequential liquid delivery
as well as for quantifying the mixing efficiency consisted
of three inlet channels for sequential delivery of three fluids to a common detection or reaction point. Red, yellow,
and blue food dyes were used to visualize timed fluid delivery. The dye solutions were introduced at opposing inlets
using an 8-channel micropipette that dispense the liquids
simultaneously. The three inlets merged to form a single
outlet channel. The outlet channel was patterned either as a
straight or in a zigzag pattern. The photomasks were made
from office paper by cutting as shown in Fig.1. The position of the liquid front was recorded with a digital SLR
camera (Nikon D3200) fitted with a macro lens (Sigma
105mm EX DG) at a rate of 50 frames per second, and
subsequently analyzed using the ImageJ software package
(Abramoff etal. 2004).

160

10 min
Paper board

140
120
100
80
60
40
20
0

10

15

20

25

30

35

Time [s]
Fig.2Water contact angle of paperboard as a function of UV-irradiation time. Reference paperboard (filled squares), paperboard coated
with TiO2 nanoparticles (X) and TiO2-coated paper after UV exposure for 5min (filled diamonds), 10min (filled triangles) and 20min
(filled circles). Inserted are pictures of a water droplet levelling 30s
after it was placed on paper with 0, 5, 10 and 20min UV exposure

13


63
Page 4 of 9

Microfluid Nanofluid (2016) 20:63

Fig.3A schematic illustrating


the mechanism for the increase
of hydroxyl group by oxidation
of carbonaceous layer

under UV irradiaon

carbonaceous layer

attributed to an increase in the number of hydroxyl groups


instigated by photocatalytic oxidation of the thin carbonaceous layer covering the TiO2 nanoparticles schematically
in Fig.3 (Teisala etal. 2013). A thorough explanation on
TiO2 photocatalytic conversion can be found in (Schneider
etal. 2014; Stepien etal. 2012, 2013; Teisala etal. 2013).
Although relatively long UV exposure times were used
here, previous work has demonstrated the same wettability conversion in a roll-to-roll process at web speeds up to
12m/min (Tuominen etal. 2014). The deposition of TiO2
nanoparticles was carried out with a pilot scale roll-to-roll
process at a speed of 50m/min. The process can easily
be scaled up to high volume production, and speeds up to
300m/min have been demonstrated. The cost of TiO2 nanoparticles deposition has been estimated at approximately
~0.09USD/m2. The cost of UV irradiation depends on the
technique used. Overall, the cost of paper-based microfluidics on a large-scale process may cost less than 0.1 USD/
m2. As a comparison, costs for photoresist SU-8 is ~$0.1
for patterning filter paper of 100cm2, using wax ~$0.01 for
patterning filter paper of 100cm2 and alkyl ketene dimer
AKD, ~$0.00001 for patterning filter paper of 100cm2 (Li
etal. 2012; Yetisen etal. 2013).
An adjacent region of high and low surface energy was
created on paper. Figure4 plots contact angle and droplet
base length as a function of time as droplet wets surfaces
with different surface energy (See also supporting information for a video). The low and high surface energy regions

Contact angle, [deg]

16

Length[mm]

100

12

80
8

60
40

20
0

10

20

30

40

50

60

Base length, [mm]

Mean CA []

120

Time, [s]
Fig.4Contact angle and base length over time as a 10l droplet
spreads on a low surface energy region (irradiation time 2min) of
channel to a higher surface energy region (irradiation time 20min).
Inserted are images of the droplet at the start, before, and after it
crosses from one region to the other

13

OH

UV irradiation

TiO2

TiO2
OH

were exposed to UV for 2 and 20min, respectively. The


ability to create surfaces with desired contact angle is
advantageous because the paper can be tailored for a given
application, e.g., creation of areas with contact angle suitable for deposition functional materials (reaction zone), or
to control surface flows, which is demonstrated in the next
section.
3.2Surface wettability contrast tocontrol flow
alongthe channel
Fluid introduced at one end of a channel flows on the surface as a result of capillary forces exerted at the liquid
airsolid interface along the fluid front. The rate of fluid
flow was altered by varying the width of the channel and
by altering surface energy along the channel. The former
is addressed in the next section. The flow was controlled
by positioning a stripe (a region with lower surface energy
relative to the main channel) normal to the flow direction.
A channel with different surface energies was created by
varying the irradiation time (Fig.2). A section of the channel that caused flow delay is referred to as a stripe. The
main channel was fabricated by exposing the paper to UV
light through a mask for 20min. The stripe region was created simultaneously by limiting the irradiation time on it
to 15min by an additional mask. The chosen irradiation
times resulted in 15 and 30 water contact angles on the
channel and stripe regions, respectively. The length and the
position of the stripe along the channel were used to control the flow. Figure5 shows flow retardation as a function
of the length of the stripe positioned at 5mm from a reference point (5mm away from the channel inlet), where the
flow was considered fully developed. Figure5b quantifies
the advancement of the liquid front by piecewise power law
fit to the data. By varying the length of the stripe, different
delay times are obtained. In the absence of the stripe (uniform channel), the liquid front took approximately 0.5s to
advance 20mm. The introduction of a stripe delayed the
liquid flow for approximately 2s for 0.5mm long stripe,
and 10s and 60s for 1mm and 2mm long stripes, respectively. The delays are caused by the liquid front stalling when reaching the lower surface energy stripes. In
the absence the stripe, the distance travelled by the fluid
front shows a linear relationship on a distance versus time
plot before slowing down due the viscous resistance and
decrease in the supply of the liquid. In the presence of the

Page 5 of 9 63

Microfluid Nanofluid (2016) 20:63

stripe, the fluid front follows the same trajectory as the


uniform channel before it reaches the stripe. On reaching the stripe, the wetting rate is slowed down because of
the change in substrate wettability. In time, the fluid front
transits over the stripe and accelerates again. The recovery
of the flow speed led to a final meniscus position approximately equal to that of the channel without a barrier stripe.
This implies that stripe impacts the delivery time but not
the overall distance the liquid wets. It is worthwhile noting
the similarities between the results obtained here and the
theory by Nikolayev etal. (Nikolayev and Beysens 2002).
The authors showed that the presence of defects on the surface does not strongly change the relaxation time of a droplet but the time lost during pinning is recovered during
slip motion. This is close to what we observe experimentally with the stripe behaving as defect along the uniform
channel. The presence of the stripe causes a net delay in the
time required for the fluid front to reach the end of the main
channel.
Liquid flow can also be manipulated by changing the
position of the stripe in the channel. Figure6 shows the

flow delays when a 1-mm-wide stripe was positioned at 5,


10 and 15mm away from the reference point. These positions lead to flow delays of nearly 10, 80 and 150s. When
the stripe is positioned closer to the channel inlet, it took
shorter time for the liquid to overcome the barrier before
accelerating again. This can be due to the excess hydrostatic pressure from the initially deposited liquid near the
channel inlet. For the stripes located at 10 and 15mm away
from the inlet, the influence of hydrostatic pressure on the
flow is insignificant and only capillary and viscous forces
control the flow. Figure6b shows piecewise power law fits
to the advancing liquid front for the different stages of the
flow, before the stripe, in the stripe and after the stripe. The
influence of hydrostatic pressure is evident in the higher
exponent (exponent=0.09) for the stripe closer to the inlet
in comparison with those further away (exponent=0.02).
The increase in delay times as the stripes were positioned further in the channel can be attributed to levelling
of the droplet which lowers the driving force for the channel flow as the liquid front advances. Figure6 shows the
receding contact angle of droplets placed in channels with

y = 0.01x3.31

y = 4E-3x3.39

Distance, mm

Distance, mm

y = 3.03x3.20
10

1
0.01

0.5mm@5mm
2mm@5mm
0.1

1mm@5 mm
uniform channel
10

10

y = 6x0.12
y = 24x0.82

1
0.01

100

y = 5x0.09
0.5mm@5mm
2mm@5mm
0.1

y = 5x0.04

1mm@5 mm
uniform channel

10

100

Time, s

Time, s

Fig. 5a Flow delay caused by changing the length of a stripe of low wettability and b piecewise power law fits to the liquid front movement

y =3E-3x2.78 y =0.04 x1.20

y = 0.01x3.31

10

1mm@5 mm
1mm@10mm

Distance, mm

Distance, mm

y = 13x0.02
10

y = 9x0.02
y = 28x0.81
y = 5x0.09

1mm@15mm
1
0.01

uniform channel
1mm@10mm

uniform channel
0.1

Time, s

10

100

1
0.01

0.1

1mm@5 mm
1mm@15mm
10

100

Time, s

Fig. 6a Delay of the liquid front obtained at different stripe locations along the channel and b piecewise power law fits to the liquid front
movement

13


63
Page 6 of 9

Microfluid Nanofluid (2016) 20:63

Contact angle, A [deg]

80

undesirable for open flow construction with volatile liquids


due to potential liquid evaporation.

70
60

3.3Control bychanging the width ofthe channel

50
40
10 mm away

30
20

15 mm away
0

Time, s

Fig.7Droplet receding contact angle before it reaches the stripe


positioned 10 and 15mm away

stripes placed at 10 and 15mm from the reference point.


The droplet motion is generally determined by the balance
of a driving capillary force and opposing forces. Therefore,
the driving force arising from the unbalanced Youngs force
related to the contact angles on either side of the droplet
may intuitively be expressed as (Kooij etal. 2012):

F = (cos A cos B )
where F is the driving force, liquid surface tension, and
A and B are the contact angles at receding and advancing angles as illustrated in Fig.7 inset. The stripe closer to
reference point, located 10mm away, shows a high receding contact angle (A) because of higher volume of the liquid leading to a high driving force. In conclusion, surface
energy differences can be used to regulate flow in a specific microfluidic application by changing the thickness of
the barrier stripe as well as the location of the stripe along
the channel. Similar to flow delays created by dissolving
sugar or erodible solvent, liquid flows continuously once
passing the delay barrier. Delay times over 60s may be

2mm channel

50

2->3@1cm

2->3@2cm

The surface flow control was achieved by creating channels with expansions and constrictions. Figure8 shows
different channel designs: a 2-mm constant width channel
that acts as reference, stripes with two different constant
widths with a narrow channel connected to a wider channel at a distance 10 and 20mm lengths from the liquid
source inlet and an expansion along the channel. The position of meniscus over time is plotted for the different channels. Initially, at time t=0, the liquid meniscus position in
the channel was set as 5mm in the analysis. The constant
width channel shows a steady velocity in the liquid front
before slowing down. This decrease in velocity is attributed to finite liquid supply in addition to viscous resistance
(Songok etal. 2014). The liquid front reached a distance
40mm in nearly 5s. This can be compared to liquid flow
through the paper matrix (Grade 1 chromatography paper),
where a liquid front reached 35mm in 50s in the experiment conducted by Zhong etal. with the constant width
paper fluidic channels fabricated by wax printing (Zhong
etal. 2012). To control the flow, the channel widths were
varied. The channel width was increased by 50% (2mm
to 3mm at 10 and 20mm away from the channel inlet)
and decreased by 50% (21mm at 10mm). The results
show similar flow characteristics before the point at which
the channel width was changed. At the transition point
from narrow to a wide channel, i.e., at 5 and 15mm, the
liquid front velocity decreased significantly. In accordance
to mass conservation, increasing the width of the channel
means increasing the area thus results in decrease of liquid
front velocity. However, moving from the wider channel to

2->1 contr

2 mm uniform channel

Meniscus position, mm

2 mm

40

2-3 mm diverging channel

20
2 mm

10

30

0
20

2-3 mm diverging channel


0

0.5

1
Time, s

1.5

2 mm

3 mm

20 mm

10
0

3 mm

10 mm

2-1 mm converging channel


2 mm

Time, s

10

1 mm

15

Fig. 8a Position of the liquid front plotted as a function of time when the channel width was increased or decreased. b Channel geometries.
The inset shows results close to origin

13

Page 7 of 9 63

Microfluid Nanofluid (2016) 20:63

a narrow channel, the flow stops at approximately 4mm


after entering the narrow channel caused by resistance at
the edge of the channel, where the free surface is pinned
due to the hydrophilic/hydrophobic boundary. Further,
the reduced contact area lowers the surface tension-driven
force acting on the contact line while maintaining the droplet pressure force acting on the cross-sectional area, hence
halting the flow.
By expanding the channel, delay times less than 1s
were observed, which may not be sufficient, whereas by
constricting the channel resulted in halting the liquid flow.
Therefore, different configurations that combine a constriction and an expansion were used to meter the flow, as presented in Fig.9. The shorter constriction (1-mm-long neck)
showed a reduced velocity at the neck with comparable
velocity as the uniform channel. However, when the length
of the neck was doubled (2mm long constriction), the flow
rate depicted by the gradient of the curve greatly declined.
Interestingly, we note that the channel with constrictions
delivered the liquid the furthest. Compared to the uniform

channel in which liquid front advanced close to 40mm in


4s, channels with 1 or 2mm long contractions allowed
flows up to 50mm in 7 and 20s, respectively. It appears
that the re-acceleration force after the constriction draws
out virtually all the liquid, thereby making the position
after the constriction a new starting point. In the diverging
converging construction along the channel, the liquid was
observed to fill up the bulb (the wider part) and not to flow
into the channel. The flow stops as a result of high resistance from the narrow channel.
3.4Sequential fluid delivery andanalysis ofmixing
characteristics
In this section, we apply the fluid control strategies
described to enable sequential fluid delivery on paper surface. Figure10 shows a device that uses three 2-mm-wide
inlet channels for sequential delivery of three fluids to a
common detection or reaction zone. As shown, sequential delivery of each of these fluids is made possible by

60

Meniscus position, mm

2 mm

50

Converging-diverging channel 2-1-2

40

2 mm
1 mm

30

Converging-diverging channel 2-1-2

20

2 mm

2mm channel
1 mm long constriction
2 mm long constriction
5 mm Bulb

10
0

uniform channel

10

15

Time, s

20

25

1 mm

Diverging-converging channel 2-3-2


2 mm 3 mm

30

2 mm

5 mm

Fig.9Meniscus position for different geometry combinations

1s

1s

2s

2s

25 s

14 s

180 s

180 s

Fig.10Sequential images showing different liquid arriving at different time into the channel and the subsequent mixing of the liquids. Scale
bar of 10mm. a Shows a straight channel and b a zigzag channel

13


63
Page 8 of 9

changing the surface energy across the inlet channels.


By changing the length of the delay stripes having lower
wettability, the fluids flow to the outlet channel at different times, first yellow, followed by red and finally blue.
Sequential delivery of liquid and by extension fluid control
may find application in for example rehydrating dried reagents before the test or in signal amplification after reaction as demonstrated by Toley etal. (2013). Sequential fluid
delivery was achieved; however, incomplete mixing was
apparent at the detection zone, indicated by the green color
encapsulated in orange ring (top right subfigure in Fig.10).
By changing the outlet channel from a straight one to a zigzag one, mixing of the fluid can be improved as shown in
bottom right subfigure of Fig.10. The zigzag outlet was
constructed to have alternating wider and narrow channel
widths. This leads to repeated acceleration and deceleration
of the advancing liquid front which encourages backflow
and laminar recirculations along the zigzag channel and
improves mixing. Mengeaud etal. (2002) who performed
mixing simulations of species in a zigzag channel concluded that the liquid recirculation induces the transversal
component of velocity that improves the mixing efficiency
(Mengeaud etal. 2002).

4Conclusion
In this work, we report two methods with which a flow
delay can be achieved in surface flow-based paper-based
fluidic devices without additional fabrication steps or user
action requirement. The first method we have shown takes
advantage of the surface energy contrast to cause a flow
delay. The surface energy difference along a channel was
accomplished by varying the UV exposure time of TiO2coated paper. The flow retardation is achieved as the liquid
flows over a higher contact angle stripe. Parameters to control the delay include the length and position of the stripe
along the channel. This broadens the time range in which
timed fluidic delivery may be achieved. We have also
demonstrated flow delay through changing the geometry
(width) of the channel. These approaches effectively control the flow without increasing the required liquid volume.
A combination of surface wettability gradient and varying
the channel geometry has been used to show sequential
delivery can be achieved without using external source of
energy. We have also shown that homogeneous mixing can
be achieved in capillary-driven surface flow. The mixing is
fostered by the back flow and flow recirculation because
it induces transversal flow. Potential shortcomings of the
proposed liquid flow control system are related to possible interactions between the TiO2 nanoparticles and the
reagents used in a device. Furthermore, the open channels
proposed herein are not suitable for assays requiring very

13

Microfluid Nanofluid (2016) 20:63

long incubation times due to liquid evaporation and drying.


Future work will aim at demonstrating PADs utilizing onpaper fabricated microfluidics including particulate liquids
such as blood.

References
Abe K, Suzuki K, Citterio D (2008) Inkjet-printed microfluidic multianalyte chemical sensing paper. Anal Chem 80:69286934
Abe K, Kotera K, Suzuki K, Citterio D (2010) Inkjet-printed paperfluidic immuno-chemical sensing device. Anal Bioanal Chem
398:885893. doi:10.1007/s00216-010-4011-2
Abramoff MD, Magalhaes PJ, Ram SJ (2004) Image processing with
ImageJ. Biophotonics Int 11:3642
Ballerini DR, Li X, Shen W (2012) Patterned paper and alternative
material as substrates for low cost microfluidic diagnostics.
Microfluid Nanofluidics 13:769787
Carrilho E, Martinez AW, Whitesides GM (2009) Understanding
wax printing: a simple micropatterning process for paper based
microfluidics. Anal Chem 81:70917095
Cate DM, Adkins JA, Mettakoonpitak J, Henry CS (2015) Recent
developments in paper-based microfluidic devices. Anal Chem
87:1941
Chen H, Jeremy C, Anagnostopoulos C, Faghri M (2012) A fluidic
diode, valves, and a sequential-loading circuit fabricated on layered paper. Lab Chip 12:29092913
Cheng C, Martinez AW, Gong J, Mace CR, Phillips ST, Carrilho E,
Mirka KA, Whitesides GM (2010) Paper-based ELISA. Angew
Chem Int Ed 49:47714774
Dungchai W, Chailapakul O, Henry CS (2011) A low-cost, simple,
and rapid fabrication for paper-based using wax screen printing.
Analyst 136:7782
Fu E, Kauffman P, Lutz B, Yager P (2010a) Chemical signal amplification in two-dimensional paper networks. Sens Actuators B
Chem 149:325328
Fu E, Lutz B, Kauffman P, Yager P (2010b) Controlled reagent transport in disposable 2D paper networks. Lab Chip 10:918920
Fu E, Ramsey SA, Kauffman P, Lutz B, Yager P (2011) Transport
in two-dimensional paper network. Microfluid Nanofluidics
10:2935
He Y, Wu Y, Fu J, Wu W (2015) Fabrication of paper-based microfluidic analysis devices: a review. RSC Adv 5:7810978127
Hitzbleck M, Avrain L, Smekens V, Lovchik RD, Mertens P,
Delamarche E (2012) Capillary soft valves for microfluidics. Lab
Chip 12:19721978
Jahanshahi-Anbuhi S, Henry A, Leung V, Sicard C, Pennings K, Pelton R, Brennan JD, Filipe CDM (2014) Paper-based microfluidics with an erodible polymeric bridge giving controlled release
and timed flow shutoff. Lab Chip 14:229236
Kooij ES, Jansen HP, Bliznyuk O, Poelsema B, Zandvliet HJW (2012)
Directional wetting on chemically patterned substrates. Colloids
Surf A Physicochem Eng Asp 413:328333
Li X, Tian J, Nguyen NT, Shen W (2008) Paper-based microfluidic
devices by plasma treament. Anal Chem 80:91319134
Li X, Ballerini R, Shen W (2012) A perspective on paper-based
microfluidics: current status and future trends. Biomicrofluidics
6:113
Lutz B, Liang T, Fu E, Ramachandran S, Kauffman P, Yager P
(2013) Dissolvable fluidic time delays for programming multistep assays in instrument-free paper diagnostics. Lab Chip
13:28402847
Mkel JM, Aromaa M, Teisala H, Tuominen M, Stepien M, Saarinen
JJ, Toivakka M, Kuusipalo J (2011) Nanoparticle deposition

Microfluid Nanofluid (2016) 20:63


from liquid flame spray onto moving roll-to-roll paperboard
material. Aerosol Sci Technol 45:827837
Martinez AW, Phillips ST, Wiley BJ, Gupta M, Whitesides GM (2008)
FLASH: a rapid method for prototyping paper-based microfluidics devices. Lab Chip 8:21462150
Mengeaud V, Josserand J, Girault HH (2002) Mixing processes in
a zigzag microchannel: finite element simulations and optical
study. Anal Chem 74:42794286
Nikolayev VS, Beysens DA (2002) Relaxation of nonspherical sessile
drops towards equilibrium. Phys Rev E 65:18
Noh H, Philips ST (2010a) Fluidic timers for time-dependent, pointof-care assays on paper. Anal Chem 82:80718078
Noh H, Philips ST (2010b) Metering the capillary-driven flow
of fluids in paper-based microfluidic devices. Anal Chem
82:41814187
Olkkonen J, Lehtinen K, Erho T (2010) Flexographically printed fluidic structures in paper. Anal Chem 82:1024610250
Osborn J, Lutz B, Fu E, Kaumann P, Stevens DY, Yager P (2010)
Microfluidics without pumps: reinventing the T-sensor and H-filter in paper network. Lab Chip 10:26592665
Sackmann EK, Fulton AL, Beebe DJ (2014) The present and
future role of microfluidics in biomedical research. Nature
507:181189
Santhiago M, Nery EW, Santos GP, Kubota LT (2014) Microfluidic
paper-based devices for bioanalytical applications. Bioanalysis
6:89106
Schneider J, Matsuoka M, Takeuchi M, Zhang J, Horiuchi Y, Anpo
M, Bahnemann DW (2014) Understanding TiO2 photocatalysis:
mechanisms and materials. Chem Rev 114:99199986
Songjaroen T, Dungchai W, Chailapakul O, Laiwattanapaisal W
(2011) Novel, simple and low-cost alternative method for fabrication of paper-based microfluidics by wax dipping. Talanta
85:25872593
Songjaroen T, Dungchai W, Chailapakul O, Henry CS, Laiwattanapaisal W (2012) Blood separation on microfluidic paper-based analytical devices. Lab Chip 12:33923398
Songok J, Tuominen M, Teisala H, Haapanen J, Mkel JM, Kuusipalo J, Toivakka M (2014) Paper-based microfluidics: fabrication
technique and dynamics of capillary-driven surface flow. Appl
Mater Interfaces 6:2006020066

Page 9 of 9 63
Stepien M, Saarinen JJ, Teisala H, Tuominen M, Aromaa M, Kuusipalo J, Mkel JM, Toivakka M (2012) Surface chemical characterization of nanoparticle coated paperboard. Appl Surf Sci
258:31193125
Stepien M, Saarinen JJ, Teisala H, Tuominen M, Aromaa M, Haapanen J, Kuusipalo J, Mkel JM, Toivakka M (2013) ToF-SIMS
analysis of UV-switchable TiO2 nanoparticle coated paper surface. Langmuir 29:37803790
Teisala H, Tuominen M, Aromaa M, Mkel JM, Stepien M, Saarinen
JJ, Toivakka M, Kuusipalo J (2010) Development of superhydrophobic coating paperboard surface using the liquid flame spray.
Surf Coat Technol 205:436445
Teisala H, Tuominen M, Stepien M, Haapanen J, Mkel JM,
Saarinen JJ, Toivakka M, Kuusipalo J (2013) Wettability conversion on the liquid flame spray generated superhydrophobic
TiO2 nanoparticle coating on paper and board by photocatalytic
decomposition of spontaneously accumulated carbonaceous
overlayer. Cellulose 20:391408
Thuo MM, Martinez RV, Lan W, Liu X, Barber J, Atkinson MBJ,
Bandarage D, Bloch J, Whitesides GM (2014) Fabrication of
low-cost paper-based microfluidic devices by embossing or cutand-stack methods. Chem Mater 26:42304237
Toley BJ, McKenzie B, Liang T, Buser JR, Yager P, Fu E (2013)
Tunable-delay shunts for paper microfluidic devices. Anal Chem
85:1154511552
Tuominen M, Teisala H, Haapanen J, Aromaa M, Mkel JM, Stepien
M, Saarinen JJ, Toivakka M, Kuusipalo J (2014) Adjustable wetting of liquid flame spray (LFS) TiO2-nanoparticle coated board:
batch-type versus roll-to-roll stimulation methods. Nord Pulp
Pap Res J 29:271279
Von Lode P (2005) Point-of-care immunotesting: approaching the
analytical performance of central laboratory methods. Clin Biochem 38:591606
Yetisen AK, Akram MS, Lowe C (2013) Paper-based microfluidic
point-of-care diagnostic devices. Lab Chip 13:22102251
Zhong ZW, Wang ZP, Huang GXD (2012) Investigation of wax and
paper materials for the fabrication of paper-based microfluidic
devices. Microsyst Technol 18:649659

13

Vous aimerez peut-être aussi