scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Microfluidic actuation by modulation of surface stresses

22 Jan 2003-Applied Physics Letters (American Institute of Physics)-Vol. 82, Iss: 4, pp 657-659
TL;DR: In this paper, the authors demonstrate the active manipulation of nanoliter liquid samples on the surface of a glass or silicon substrate by combining chemical surface patterning with electronically addressable microheater arrays.
Abstract: We demonstrate the active manipulation of nanoliter liquid samples on the surface of a glass or silicon substrate by combining chemical surface patterning with electronically addressable microheater arrays. Hydrophilic lanes designate the possible routes for liquid migration while activation of specific heater elements determine the trajectories. The induced temperature fields spatially modulate the liquid surface tension thereby providing electronic control over the direction, timing, and flow rate of continuous streams or discrete drops. Temperature maps can be programed to move, split, trap, and mix ultrasmall volumes without mechanically moving parts and with low operating voltages of 2–3 V. This method of fluidic actuation allows direct accessibility to liquid samples for handling and diagnostic purposes and provides an attractive platform for palm-sized and battery-powered analysis and synthesis.

Summary (3 min read)

Introduction

  • SINCE THE eighties, the resonance method to improve thegain of printed antennas has been well documented [1]–[3].
  • Frequency selective surfaces (FSS) [13] based on the Fabry-Perot effect [14] have also been proposed as an alternative to dielectric EBGs for gain enhancement.
  • The distance between the FSS superstrate and the ground plane of the antenna, which determines the resonant frequency, still needs to be about of the resonant frequency of the resonator.

II. META-SURFACE DESIGN

  • During the recent years, several double negative materials (DNG) have been designed by different authors [23]–[25] several of them based on the topology proposed by [26], i.e., a combination of split ring resonators (SRRs) and wires.
  • The meta-surface under study is based on the unit cell proposed by Prof. Ziolkowski in [25].
  • The material selected to fabricate the layers was RT/Duroid 5880, a low loss dielectric characterized by the parameters , loss tangent , thickness 0.254 mm and copper cladding 70 on both faces.
  • The transmission and reflection properties of the manufactured meta-surface were tested under waveguide excitation.
  • The measured results are depicted in Fig.

III. RADIATION PERFORMANCE

  • As was mentioned before, this meta-surface is basically a resonant structure exhibiting pass and stop bands.
  • It is formed by a finite periodic repetition of the unit cell.
  • By tuning a dipole antenna to the pass band frequency of the superstrate, an in-phase resonance of the unit cells will be induced leading to a more uniform illumination of the superstrate.
  • The dimensions and position of the dipole were optimized by simulating it embedded in a dielectric slab of and loaded with the superstrate.
  • Fig. 4 shows the deviation of the resonant frequency of the dipole with and without superstrate with respect to the optimized configuration when the dimensions of the dipole are slightly modified.

A. Parameter

  • The parameter of the dipole with the superstrate was measured by using a network analyzer (Marconi 6210 Reflection Analyzer).
  • Based on the simulation optimization process the dipole was placed just above the superstrate.
  • The set-up for the measurements is shown in Fig.
  • The influence of the meta-surface in the impedance matching of the dipole was analyzed and measured by varying the number of periods of the superstrate in the x direction from 4 to 12.
  • As expected, an impedance matching better than 12 dB with a resonant frequency around 11.1 GHz and a deviation smaller than 3.5% with respect to the central frequency has been obtained.

B. Resonant Frequency

  • Fig. 7 shows the dependence of the resonant frequency (minimum ) of the configuration (meta-surface + dipole) and the frequency of maximum gain at boresight with the number of cells of the superstrate.
  • The process of deriving the gain will be explained in the next section.
  • It can be observed that both parameters do not exactly coincide, but there is a slight frequency shift between them.
  • A similar tendency was obtained in both cases; as the number of cells increases, the resonant frequency decreases varying from 11.19 GHz in the case of 4 cells to 10.9 GHz in the case of 12 cells [this agrees with the result obtained with the waveguide measurements (see Fig. 2)].
  • The resonant frequency curve flattens for larger superstrates since the meta-surface trends to behave as a more uniform media resonating at 10.9 GHz instead of a set of single scatterers.

C. Gain

  • In order to characterize the different configurations in terms of absolute-gain, the two-antenna method described in [27] was followed.
  • The gain of the radiating configurations was measured in an anechoic chamber and compared with the one of a single dipole by using a horn antenna as receiver and the dipole (with and without superstrate) as transmitter.
  • The superstrate was placed close to the dipole, as in the case of the measurements, in order to maximize the power radiated at boresight.
  • The set-up used for these measurements is shown in Fig.
  • The reason for this is the appearance of a standing wave in the transversal direction of the configuration, which limits the maximum size of the structure that achieves large gain.

D. Radiation Pattern

  • Once the frequency of maximum gain was known (see Fig. 9), the radiation patterns were measured at that frequency.
  • In order to avoid reflections from the clamp and the feeding cable, the back side of the antenna was covered with the absorbing material shown in Fig. 8(b).
  • The back radiation pattern (from 90 to 270 ) was not measured because of physical limitations in the test set-up (see the back side feeding network and clamp in Fig. 8).
  • Fig. 11 shows the H and E-plane radiation patterns for the case of nine cells.
  • The back radiation level at 180 was around 5 dB what means a front-back radiation of 10 dB.

E. Radiation Efficiency

  • The radiation efficiency is defined as the ratio of the total power radiated by the antenna to the total power accepted by the antenna at its terminals during radiation.
  • Following this approach, the radiation efficiency has been computed for all the configurations, (see Fig. 13 solid line), taking into account 1 dB errors in the gain measurements.
  • For superstrates smaller or larger, the radiation efficiency reduces, but it always exceeds 50%.
  • So, the real part of the input impedance with the sphere in place will be .
  • Thus by making two impedance measurements, without and with the cap, the radiation efficiency can be determined by using (5).

F. Effective Area

  • Once the maximum directivity at boresight of the configuration is obtained, the effective area of the meta-surface can be calculated by applying (6) (6) Fig. 14 shows a comparison between the effective area derived from the measurements (taking into account the errors in the measurements of the gain) and the physical area of the meta-surface.
  • When the superstrate is small (4 unit cells; equal to ), it can be observed that the effective area is larger than the physical one.
  • This is an artefact effect produced by the mathematical definition of the effective area when the radiating surface is not wide enough and there are not metallic walls surrounding the antenna.
  • This is a known phenomena described by Balanis for dipole antennas [27] p. 83.
  • For larger superstrates (from five to eight cells), the effective area is similar to the physical one, what means that the meta-surface is completely illuminated, i.e., a uniform illumination has been achieved.

IV. MULTIFREQUENCY ANTENNA CONFIGURATION

  • The results obtained in the previous sections show that the physical area of the proposed structure can be used very effectively in order to enhance the radiation performance of planar antennas due to the uniform illumination of the superstrate.
  • It is possible to place the same type of unit cell but with a different resonant frequency (same shape but different size) on both sides.
  • In order to implement it, the unit cell explained in Section II has been scaled to be working at a higher frequency, around 12.5 GHz.
  • By combining superstrates consisting of low resonant frequency (LRF) unit cells and high resonant frequency (HRF) ones, and tuning dipoles to these frequencies, the MFAA is designed.
  • The MFAA studied is schematically depicted in Fig. 16.

V. CONCLUSION

  • A novel implementation to enhance the radiation performances of a dipole antenna based on the use of meta-surfaces as superstrate has been presented.
  • A finite uniform meta-surface has been characterized in terms of parameter, gain, radiation pattern, radiation efficiency and effective area.
  • Measurements of different configurations have proven an enhancement of the gain at boresight of about 3.5 1 dB with a reduction of the H-plane endfire radiation of about 15 dB.
  • For larger superstrates, the effective area does not increase since the maximum area that the dipole can illuminate has been estimated around .
  • Based on this configuration, a compact multifrequency antenna array has been simulated, showing the gain enhancement and low coupling between elements.

Did you find this useful? Give us your feedback

Content maybe subject to copyright    Report

Microfluidic actuation by modulation of surface stresses
Anton A. Darhuber, Joseph P. Valentino, Jeffrey M. Davis, and Sandra M. Troian
a)
Microfluidic Research and Engineering Laboratory, Department of Chemical Engineering,
Princeton University, Princeton, New Jersey 08544
Sigurd Wagner
Department of Electrical Engineering, Princeton University, Princeton, New Jersey 08544
Received 26 June 2002; accepted 21 November 2002
We demonstrate the active manipulation of nanoliter liquid samples on the surface of a glass or
silicon substrate by combining chemical surface patterning with electronically addressable
microheater arrays. Hydrophilic lanes designate the possible routes for liquid migration while
activation of specific heater elements determines the trajectories. The induced temperature fields
spatially modulate the liquid surface tension thereby providing electronic control over the direction,
timing, and flow rate of continuous streams or discrete drops. Temperature maps can be programed
to move, split, trap, and mix ultrasmall volumes without mechanically moving parts and with low
operating voltages of 23 V. This method of fluidic actuation allows direct accessibility to liquid
samples for handling and diagnostic purposes and provides an attractive platform for palm-sized and
battery-powered analysis and synthesis. © 2003 American Institute of Physics.
DOI: 10.1063/1.1537512
Miniaturized automated systems for liquid routing, mix-
ing, analysis, and synthesis are rapidly expanding diagnostic
capabilities in medicine, genomic research, and material
science.
1
Liquid flow in microchannels can be regulated by
pressure gradients,
2
thermocapillary pumping,
3
electrokinetic
forces,
4,5
or magnetohydrodynamic pumping.
6,7
Electrowetting
8,9
and dielectrophoresis
10
have also been used
to move droplets on an open surface. These techniques typi-
cally require high operating voltage and high electrolyte con-
centrations. We demonstrate a different method for liquid
handling and transport that uses programmable surface tem-
perature distributions, in conjunction with chemical substrate
patterning, to provide electronic control over the direction,
timing, and flow rate. This method capitalizes on the large
surface-to-volume ratio inherent in microscale systems since
the gasliquid and liquidsolid surface energies are modu-
lated to induce and confine flow. It works equally well with
polar or nonpolar liquids, requires no mechanically moving
parts and operates at very low voltages. The open architec-
ture is best suited to liquids of low volatility. Encapsulation
schemes that retain the free liquidair interfaces can mini-
mize evaporative loss.
Local heating of a liquid film at a position x reduces the
surface tension
(x) to produce a thermocapillary shear
stress
˜
nˆ
(
/
T) T that pulls liquid toward re-
gions of cooler surface temperature T.
11–13
Since
/
T is
essentially temperature independent for all liquids, the driv-
ing force and flow direction are proportional to T. For a
thin flat liquid layer, the average flow speed and flow rate
per unit width are given by
(x) h(x,t)
˜
nˆ /2
(x) and
Q(x,t) h(x,t)
(x,t), where h(x,t) is the film thickness
and
(x) the local viscosity. This phenomenon forms the
basis of our microfluidic device for actuating continuous
streams and discrete droplets.
The sample layout
14
for driving continuous streams is
shown in Fig. 1a. Hydrophilic stripes connect pairs of 4.5-
mm-wide square reservoir pads; the gray areas are chemi-
cally treated to repel the liquid.
15,16
A constant temperature
gradient dT/dx is generated by an embedded heating resistor
vertical black line on the left and a cooling block dotted
rectangle on the right. Figure 1b depicts a microstream
emerging from a reservoir into a hydrophilic stripe subject to
a streamwise temperature gradient. The liquids used were
polydimethyl-siloxane PDMS,
5 and 20 mPa s,
20.6 mN/m, and d
/dT⫽⫺0.060 mN/m °C at 25 °C and
dodecane (
1.35 mPa s,
25.5 mN/m, and d
/dT
⫽⫺0.091 mN/m °C at 25 °C.
The speed of a continuous microstream depends on its
length L, channelwidth w, dT/dx, and the deposited volume
V
s
. We have obtained a flow speed of 600
m/s for PDMS
(
5mPasat25°C with dT/dx 1.44 °C/mm, w
800
m, and V
s
8
l. Flow speeds higher than 1 mm/s
can be obtained with smaller
, larger
/
T, and narrower,
continuously fed reservoirs. Figure 2a shows the position
a
Author to whom correspondence should be addressed; electronic mail:
stroian@princeton.edu
FIG. 1. a Layout of a chemically patterned silicon sample with uniform
temperature gradient used for moving continuous streams. Gray areas are
hydrophobic, white areas hydrophilic. Liquid 2–8
l deposited on a
diamond-shaped reservoir pad is propelled on a microstripe width 100800
m due to the thermocapillary stress. The black rectangles are electrical
contact pads connecting to the subsurface heating resistor black line.The
dashed rectangle denotes the cooling block. b Top-view optical micrograph
of a rivulet moving on a 300-
m-wide channel.
APPLIED PHYSICS LETTERS VOLUME 82, NUMBER 4 27 JANUARY 2003
6570003-6951/2003/82(4)/657/3/$20.00 © 2003 American Institute of Physics
Downloaded 15 Sep 2006 to 131.215.240.9. Redistribution subject to AIP license or copyright, see http://apl.aip.org/apl/copyright.jsp

dependence of the front speed for PDMS (
20 mPa s) and
four different gradients. The solid lines represent numerical
simulations of the front speed obtained from a lubrication
equation for the evolution of centerline film thickness h(x,t)
according to
h
t
2
h
2
5
64h
3
315
h
xx
x
192h
3
105
w
2
h
x
x
0, 1
where subscripts denote partial differentiation. The film
shape and speed are determined by the thermocapillary driv-
ing stress term 2, capillary forces due to streamwise and
lateral surface curvature terms 3 and 4, and the
temperature-dependent viscosity.
15
The agreement with ex-
periment is excellent. The observed decrease in speed shown
in Fig. 2a is caused by the decrease in capillary forces
16
and the increase in viscosity with distance. The relevant di-
mensionless numbers based on parameter values at the inlet
x0 are (l/w)
2
, where l (
h
3
/3
v
)
1/3
is the dynamic cap-
illary length,
17
and
/p
0
where p
0
is the reservoir pressure.
15
For
/p
0
1, the inlet film height h
0
is controlled by p
0
,
which is essentially proportional to V
s
.
16
A scaling
analysis
15
then gives
w
2
V
s
/
. In the opposite limit
/p
0
1,
determines the inlet height
15
and the front speed
scales as
(w
)
3/2
/
. Figure 2b shows experimental data
and least-square fits for the front speed as a function of w at
x 18 mm. To a good approximation, the data follow a
power law
w
, where the exponents
confirm the trend
from 2 to 3/2. The dependence of the front speed on dT/dx
shown in Fig. 2c also follows the scaling predictions
v
, where
1 and 1.47. The two data sets illustrate the
transition from reservoir to shear-dominated regimes. For in-
termediate values of
/p
0
, the flow speed does not conform
to power laws but can only be determined by full numerical
simulations.
15
Mixing poses a difficulty in microfluidic devices due to
the absence of turbulent convection. Figures 3a–3c depict
computed convection patterns obtained with a purely trans-
verse thermal gradient dT/dy in the absence of diffusion.
The interfacial area A(t) between the two liquids increases at
an essentially constant rate by a factor of 150 in just 30 s.
Simulations including molecular diffusion with coefficient
D 10
12
m
2
/s show a reduction in the mixing time t
mix
by
three orders of magnitude compared to purely diffusive mix-
ing. The presence of both convection and diffusion leads to a
nonmonotonic scaling of t
mix
with D similar to TaylorAris
dispersion.
1820
Since thermocapillary flow is proportional to
h(x,t), even faster mixing can be obtained by directing the
flow onto a wider hydrophilic patch that supports thicker
films. By imposing an additional thermal gradient dT/dx
parallel to the microstripe, the streamlines follow helical
paths thereby significantly increasing the interfacial area be-
tween merging streams. Figure 3d presents experimental
flow patterns on a hydrophilic stripe inclined by 45° to the
direction of the thermal gradient. The solid symbols desig-
nate the measured trajectories
x(t),y(t)
of tracer particles
convected with the flow. A larger spacing between subse-
quent points signifies a faster flow near the airliquid
interface—a smaller spacing indicates slower flow near the
solid substrate.
The resistor and channel layout used for actuation of
FIG. 2. Front speed of continuous streams. a Position dependence for four
temperature gradients on a 500-
m-wide stripe. Symbols: experimental
data; solid lines: simulations. The film thickness at the stripe inlet was
treated as a fitting parameter (h
0
402
m). b Dependence on w for
four thermal gradients. c Dependence on dT/dx for w 300
m.
FIG. 3. ac Mixing of two liquids black and white by thermocapillary
convection induced by a transverse temperature gradient. The stripe is 500
m wide, h
0
20
m, dT/dy4 °C/mm, and
20 mPa s. The curves
shown in b represent the streamlines of the convective flow. d Experi-
mental positions
x(t),y(t)
of 5-
m-diam tracer particles in a rivulet flow-
ing at a 45° angle from the direction of T (w 530
mand
18 mPa s). The time increment between subsequent data points is t
6 s. The gray band signifies a portion of the slanted hydrophilic stripe.
658 Appl. Phys. Lett., Vol. 82, No. 4, 27 January 2003 Darhuber
et al.
Downloaded 15 Sep 2006 to 131.215.240.9. Redistribution subject to AIP license or copyright, see http://apl.aip.org/apl/copyright.jsp

discrete drops is shown in Fig. 4. The surface temperature is
electronically programmed through an array of embedded
heating resistors light gray. Discrete drops or continuous
streams can be moved along hydrophilic pathways orange
and directed around corners. Local temperature control al-
lows splitting of a liquid stream into droplets of predefined
volume. Figures 4a–4e show the generation of a doublet
from a single drop. The rate of film thinning above a heated
resistor is essentially proportional to the heating power, but
independent of w and the film thickness.
21
The rupture of the
small connecting thread leaves behind satellite droplets,
which can be absorbed by dispatching one of the main drops
to collect the residue. The electrical power used to split and
propel drops of dodecane was typically less than 40 mW per
resistor. The power can be reduced by at least one order of
magnitude by using a polymeric substrate of low thermal
conductivity instead of glass.
22
Battery operation of this mi-
crofluidic device is, therefore, practicable.
Figures 4f–4l depict electronic propulsion of discrete
droplets along partially wetting microstripes. The heating el-
ements are slightly narrower near the microstripe edges.
These constrictions increase the local resistive heating,
which causes a slightly elevated temperature at the stripe
edges. This helps center the streaming drops and films later-
ally. Figures 4f–4i show a dodecane drop moving from
right to left through an intersection of two partially wetting,
1000-
m-wide stripes on a glass substrate. Figures 4j–4l
show a drop being pulled around a corner. The average drop-
let speed
d
is controlled by the electronic switching
interval—in Figs. 4f–4i
d
56
m/s and in Figs. 4j
4l,
v
d
26
m/s. In these series, the applied voltage was
2.36 V and the heater resistance 140 . The control algo-
rithm involved sequentially powering one resistor ahead of
the drop 20 mW and two resistors behind it 共⭐40 mW each
in a ramp-like fashion. The drop position is, therefore,
known to a precision equal to the resistor width, which is
advantageous for the simultaneous routing of a multitude of
droplets. On completely wetting surfaces, a thin residue film
can trail a moving droplet since the receding contact angle is
vanishingly small. Such residual films, which can cross con-
taminate liquid samples, can be eliminated, by making the
surface pathways partially wetting.
We have demonstrated that electronically addressable
heating arrays coupled with chemical surface patterning can
be used to actuate and manipulate continuous streams or dis-
crete drops of liquid on a solid substrate. Numerical simula-
tions for the flow speed of continuous microstreams are in
excellent agreement with experiment. Electronic control en-
ables transport with preset velocity, efficient mixing of sta-
tionary or moving liquids, and merging or splitting of liquids
into specified volumes. Temperature maps can be pro-
grammed to trap drops at a specified location for analysis
and synthesis. As the presence of an overlying liquid film
affects the thermal response time of the resistors, they can be
used as position sensors providing feedback control.
22
This
device can also serve as a chemical sensor since the liquid is
in continuous contact with the ambient vapor phase. The
variety of tasks that can be accomplished with high-
resolution thermal maps should, therefore, inspire alternative
designs for pocket-sized diagnostic devices.
This work is supported by NSF grant CTS-0088774,
MRSEC grant DMR-9809483, a NJCST grant, and an ND-
SEG fellowship J.M.D..
1
D. Meldrum, Genome Res. 10, 1288 2000.
2
P. Gravesen, J. Branebjerg, and O. S. Jensen, J. Micromech. Microeng. 3,
168 1993.
3
T. S. Sammarco and M. A. Burns, AIChE J. 45,3501999.
4
J. M. Ramsey, S. C. Jacobson, and M. R. Knapp, Nat. Med. N.Y. 1, 1093
1995.
5
G. H. W. Sanders and A. Manz, Trends Anal. Chem. 19,3642000.
6
W. Ritchie, Philos. Trans. R. Soc. London 122, 279 1832.
7
J. Jang and S. S. Lee, Sens. Actuators A 80,842000.
8
J. Lee and C.-J. Kim, J. Microelectromech. Syst. 9, 171 2000.
9
M. G. Pollack, R. B. Fair, and A. D. Shenderov, Appl. Phys. Lett. 77,1725
2000.
10
T. B. Jones, M. Gunji, M. Washizu, and M. J. Feldman, J. Appl. Phys. 89,
1441 2001.
11
L. G. Leal, Laminar and Convective Transport Processes Butterworth-
Heinemann, Boston, 1992.
12
V. G. Levich, Physicochemical Hydrodynamics Prentice-Hall, Englewood
Cliffs, NJ, 1962.
13
V. Ludviksson and E. N. Lightfoot, AIChE J. 17, 1166 1971.
14
Si wafers were first coated with 200 nm SiN
x
and 200 nm SiO
2
, deposited
by plasma enhanced chemical vapor deposition PECVD at 250 °C. A
100 nm Au layer was then deposited by an electron beam evaporation
EBE and patterned into heating elements using a lift-off technique. The
heating elements were passivated with 600 nm PECVD SiO
2
, on top of
which the chemical pattern was defined in a layer of perfluorooctyl-
trichlorosilane PFOTS by photolithography. Ti heaters and Au contacts
were deposited on Corning 1737 glass slides by EBE. The hydrophilic
stripes were defined on a 200 nm PECVD SiO
2
layer using PFOTS.
15
A. A. Darhuber, J. M. Davis, S. M. Troian, and W. Reisner, Phys. Fluids.
submitted.
16
A. A. Darhuber, S. M. Troian, and W. Reisner, Phys. Rev. E 64,1603
2001.
17
S. M. Troian, E. Herbolzheimer, S. A. Safran, and J. F. Joanny, Europhys.
Lett. 10,251989.
18
G. I. Taylor, Proc. R. Soc. London, Ser. A 219, 186 1953.
19
R. Aris, Proc. R. Soc. London, Ser. A 235,671956.
20
A. A. Darhuber, J.-Z. Chen, J. M. Davis, and S. M. Troian unpublished.
21
A. A. Darhuber and S. M. Troian unpublished.
22
A. A. Darhuber, S. M. Troian, and S. Wagner, J. Appl. Phys. 91,5686
2002.
FIG. 4. ae Thermally induced splitting of a dodecane drop on a par-
tially wetting stripe (w 1000
m). The voltage applied to the microheater
155 ⍀兲 was 2.5 V. The images were recorded at t 0, 6.0, 7.5, 8.0, and 8.5
s. fi Dodecane drop propelled through an intersection outlined by the
dark gray pattern (w 1000
m, time lapse 104 s. jl Dodecane drop
turning a 90° corner time lapse 164 s.
659Appl. Phys. Lett., Vol. 82, No. 4, 27 January 2003 Darhuber
et al.
Downloaded 15 Sep 2006 to 131.215.240.9. Redistribution subject to AIP license or copyright, see http://apl.aip.org/apl/copyright.jsp
Citations
More filters
Journal ArticleDOI
TL;DR: A review of the physics of small volumes (nanoliters) of fluids is presented, as parametrized by a series of dimensionless numbers expressing the relative importance of various physical phenomena as mentioned in this paper.
Abstract: Microfabricated integrated circuits revolutionized computation by vastly reducing the space, labor, and time required for calculations. Microfluidic systems hold similar promise for the large-scale automation of chemistry and biology, suggesting the possibility of numerous experiments performed rapidly and in parallel, while consuming little reagent. While it is too early to tell whether such a vision will be realized, significant progress has been achieved, and various applications of significant scientific and practical interest have been developed. Here a review of the physics of small volumes (nanoliters) of fluids is presented, as parametrized by a series of dimensionless numbers expressing the relative importance of various physical phenomena. Specifically, this review explores the Reynolds number Re, addressing inertial effects; the Peclet number Pe, which concerns convective and diffusive transport; the capillary number Ca expressing the importance of interfacial tension; the Deborah, Weissenberg, and elasticity numbers De, Wi, and El, describing elastic effects due to deformable microstructural elements like polymers; the Grashof and Rayleigh numbers Gr and Ra, describing density-driven flows; and the Knudsen number, describing the importance of noncontinuum molecular effects. Furthermore, the long-range nature of viscous flows and the small device dimensions inherent in microfluidics mean that the influence of boundaries is typically significant. A variety of strategies have been developed to manipulate fluids by exploiting boundary effects; among these are electrokinetic effects, acoustic streaming, and fluid-structure interactions. The goal is to describe the physics behind the rich variety of fluid phenomena occurring on the nanoliter scale using simple scaling arguments, with the hopes of developing an intuitive sense for this occasionally counterintuitive world.

4,044 citations

Journal ArticleDOI
TL;DR: An overview of flows in microdevices with focus on electrokinetics, mixing and dispersion, and multiphase flows is provided, highlighting topics important for the description of the fluid dynamics: driving forces, geometry, and the chemical characteristics of surfaces.
Abstract: Microfluidic devices for manipulating fluids are widespread and finding uses in many scientific and industrial contexts. Their design often requires unusual geometries and the interplay of multiple physical effects such as pressure gradients, electrokinetics, and capillarity. These circumstances lead to interesting variants of well-studied fluid dynamical problems and some new fluid responses. We provide an overview of flows in microdevices with focus on electrokinetics, mixing and dispersion, and multiphase flows. We highlight topics important for the description of the fluid dynamics: driving forces, geometry, and the chemical characteristics of surfaces.

3,307 citations


Cites background from "Microfluidic actuation by modulatio..."

  • ...Two distinct configurations can be distinguished: drop formation and movement in small channels, which we discuss in Sections 4.2 and 4.3, and drop movements on substrates using micron-scale control of interfacial energies, which we only briefly mention (e.g., Darhuber et al. 2003)....

    [...]

Journal ArticleDOI
TL;DR: In this paper, the principles underlying common techniques for actuation of droplets and films on homogeneous, chemically patterned, and topologically textured surfaces by modulation of normal or shear stresses are reviewed.
Abstract: Development and optimization of multifunctional devices for fluidic manipulation of films, drops, and bubbles require detailed understanding of interfacial phenomena and microhydrodynamic flows Systems are distinguished by a large surface to volume ratio and flow at small Reynolds, capillary, and Bond numbers are strongly influenced by boundary effects and therefore amenable to control by a variety of surface treatments and surface forces We review the principles underlying common techniques for actuation of droplets and films on homogeneous, chemically patterned, and topologically textured surfaces by modulation of normal or shear stresses

474 citations


Cites background from "Microfluidic actuation by modulatio..."

  • ...2004; Darhuber et al. 2003b,c; Ford & Nadim 1994; Smith 1995; Yarin et al. 2002)....

    [...]

  • ...Inlet position at x = 0 (Darhuber et al. 2003a)....

    [...]

  • ...Darhuber et al. (2003a) examined the horizontal thermocapillary spreading of Newtonian liquids emanating from a small reservoir pad onto a wettable microstripe subject to a linear A nn u....

    [...]

  • ...…scale, , which determines the longitudinal extent of capillary deformation near the moving front for an infinitely wide film, and the pattern feature size w. Experimental measurements for a wide parameter range show excellent agreement with numerical solutions of Equation 3 (Darhuber et al. 2003a)....

    [...]

  • ...(j–l) Dodecane droplet turning a 90◦ corner ( ttotal = 164 s) (Darhuber et al. 2003b)....

    [...]

Journal ArticleDOI
TL;DR: It is reported that liquids perform self-propelled motion when they are placed in contact with hot surfaces with asymmetric (ratchetlike) topology and proposed that liquid motion is driven by a viscous force exerted by vapor flow between the solid and the liquid.
Abstract: We report that liquids perform self-propelled motion when they are placed in contact with hot surfaces with asymmetric (ratchetlike) topology. The pumping effect is observed when the liquid is in the Leidenfrost regime (the film-boiling regime), for many liquids and over a wide temperature range. We propose that liquid motion is driven by a viscous force exerted by vapor flow between the solid and the liquid.

466 citations

Journal ArticleDOI
TL;DR: In this paper, a method of transferring and stacking metal layers onto a polydimethylsiloxane (PDMS) substrate by using serial and selective etching techniques was proposed.
Abstract: In recent years, there has been considerable progress on fabricating microfluidic devices with multiple functionalities, with the goal of attaining lab-on-a-chip [1–3] integration. These efforts have benefited from the development of microfabrication technologies such as soft lithography. [4] In this context the material polydimethylsiloxane (PDMS) has played an important role, not only serving as the stamp for pattern transfer, but also as an unique material in chip fabrication owing to its properties such as transparency, biocompatibility, and good flexibility. [5] Because such microfluidic devices may be constructed using simple manufacturing techniques such as micromolding, they are generally inexpensive to produce. By employing PDMS, micropumps, valves, mixer/reactors, and other components have been integrated into all-in-one chips with complex functionalities, used in chemical reactions, bio-analysis, drug discovery, etc. [2] However, PDMS is a non-conducting polymer, on which patterning metallic structures during the fabrication of microdevices is challenging due to the weak adhesion between the metal and PDMS. Hence, the integration of conducting structures into bulk PDMS has been a critical issue, especially for those applications such as electrokinetic micropumps, microsensors, microheaters, electro-rheological (ER) actuators, etc., [6–8] which require electrodes for control and signal detection. Patterning metallic structures is popular in microelectronics, but the metals cannot adhere to PDMS strongly due to the low surface energy of PDMS. Lee et al. reported the transfer and subsequent embedding of thin films of gold patterns into PDMS via chemical adhesion mediated by a silane coupling agent. [9] Lim et al. [10] developed a method of transferring and stacking metal layers onto a PDMS substrate by using serial and selective etching techniques. However, the incompatibility between PDMS and the metal usually caused failures in the fabrication process, especially in the bonding of thin layers. To minimize the difference in material properties, other conductive materials were considered. Gawron et al. reported the embedding of thin carbon fibers into PDMS-based microchips for capillary electrophoresis detection. [11] Carbon

381 citations

References
More filters
Journal ArticleDOI
TL;DR: In this paper, it was shown analytically that the distribution of concentration produced in this way is centred on a point which moves with the mean speed of flow and is symmetrical about it in spite of the asymmetry of the flow.
Abstract: When a soluble substance is introduced into a fluid flowing slowly through a small-bore tube it spreads out under the combined action of molecular diffusion and the variation of velocity over the cross-section. It is shown analytically that the distribution of concentration produced in this way is centred on a point which moves with the mean speed of flow and is symmetrical about it in spite of the asymmetry of the flow. The dispersion along the tube is governed by a virtual coefficient of diffusivity which can be calculated from observed distributions of concentration. Since the analysis relates the longitudinal diffusivity to the coefficient of molecular diffusion, observations of concentration along a tube provide a new method for measuring diffusion coefficients. The coefficient so obtained was found, with potassium permanganate, to agree with that measured in other ways. The results may be useful to physiologists who may wish to know how a soluble salt is dispersed in blood streams.

4,530 citations

Journal ArticleDOI
TL;DR: In this paper, it was shown that the rate of growth of the variance is proportional to the sum of the molecular diffusion coefficient and the Taylor diffusion coefficient, where U is the mean velocity and a is a dimension characteristic of the cross-section of the tube.
Abstract: Sir Geoffrey Taylor has recently discussed the dispersion of a solute under the simultaneous action of molecular diffusion and variation of the velocity of the solvent. A new basis for his analysis is presented here which removes the restrictions imposed on some of the parameters at the expense of describing the distribution of solute in terms of its moments in the direction of flow. It is shown that the rate of growth of the variance is proportional to the sum of the molecular diffusion coefficient, D, and the Taylor diffusion coefficient $\kappa $a$^{2}$U$^{2}$/D, where U is the mean velocity and a is a dimension characteristic of the cross-section of the tube. An expression for $\kappa $ is given in the most general case, and it is shown that a finite distribution of solute tends to become normally distributed.

2,334 citations

Journal ArticleDOI
TL;DR: In this article, a microactuator for rapid manipulation of discrete microdroplets is presented, which is accomplished by direct electrical control of the surface tension through two sets of opposing planar electrodes fabricated on glass.
Abstract: A microactuator for rapid manipulation of discrete microdroplets is presented. Microactuation is accomplished by direct electrical control of the surface tension through two sets of opposing planar electrodes fabricated on glass. A prototype device consisting of a linear array of seven electrodes at 1.5 mm pitch was fabricated and tested. Droplets (0.7–1.0 μl) of 100 mM KCl solution were successfully transferred between adjacent electrodes at voltages of 40–80 V. Repeatable transport of droplets at electrode switching rates of up to 20 Hz and average velocities of 30 mm/s have been demonstrated. This speed represents a nearly 100-fold increase over previously demonstrated electrical methods for the transport of droplets on solid surfaces.

1,471 citations

Journal ArticleDOI
TL;DR: An overview of research activities in the field of fluid components or systems built with microfabrication technologies is given in this paper, focusing on the fluidic behaviour of the various devices, such as valves, pumps and flow sensors as well as the possibilities and pitfalls related to the modelling of these devices using simple flow theory.
Abstract: An overview is given of research activities in the field of fluid components or systems built with microfabrication technologies. This review focuses on the fluidic behaviour of the various devices, such as valves, pumps and flow sensors as well as the possibilities and pitfalls related to the modelling of these devices using simple flow theory. Finally, a number of microfluidic systems are described and comments on future trends are given.

1,153 citations

Journal ArticleDOI
TL;DR: In this article, the authors demonstrate a new class of high-speed DEP actuators, including wallless flowstructures, siphons, and nanodroplet dispensers that operate with water.
Abstract: Water, like any polarizable medium, responds to a nonuniform electric field by collecting preferentially in regions of maximum field intensity. This manifestation of dielectrophoresis(DEP) makes possible a variety of microelectromechanicalliquid actuation schemes. In particular, we demonstrate a new class of high-speed DEP actuators, including “wall-less” flowstructures, siphons, and nanodroplet dispensers that operate with water. Liquid in these microfluidic devices rests on a thin, insulating, polyimide layer that covers the coplanar electrodes. Microliter volumes of water, deposited on these substrates from a micropipette, are manipulated, transported, and subdivided into droplets as small as ∼7 nl by sequences of voltage application and appropriate changes of electrode connections. The finite conductivity of the water and the capacitance of the dielectric layer covering the electrodes necessitate use of rf voltage above ∼60 kHz. A simple RC circuit model explains this frequency-dependent behavior. DEP actuation of small water volumes is very fast. We observe droplet formation in less than 0.1 s and transient, voltage-driven movement of water fingers at speeds exceeding 5 cm/s. Such speed suggests that actuation can be accomplished using preprogrammed, short applications of the rf voltage to minimize Joule heating.

443 citations

Frequently Asked Questions (1)
Q1. What contributions have the authors mentioned in the paper "Microfluidic actuation by modulation of surface stresses" ?

The authors demonstrate the active manipulation of nanoliter liquid samples on the surface of a glass or silicon substrate by combining chemical surface patterning with electronically addressable microheater arrays. This method of fluidic actuation allows direct accessibility to liquid samples for handling and diagnostic purposes and provides an attractive platform for palm-sized and battery-powered analysis and synthesis.