You are on page 1of 13

Acoustic Microfluidics

Samuel R. Wilton

Acoustic wave applications within microfluidics have made an enormous impact on the
state of the industry and continue to be imperative to the evolution of fully integrated lab-
on-a-chip devices. Innovations utilizing acoustic wave phenomena to manipulate
particles and fluids have allowed scientists to perform functions faster, more easily, more
efficiently, with less power consumption, and in certain applications, better than any
other methods. Various applications of acoustics to manipulate particle flow and
alignment, mix low Reynolds number microfluidic solutions, and propagate fluids in a
microfluidic channel are discussed thoroughly in this review paper.

I. INTRODUCTION

Piezoelectric transducers [1], devices capable of converting electrical signals into


acoustic pressure waves, allow a broad band of acoustical applications and innovations to
funnel into microfluidics technology. Particle manipulation and sorting, which may
require specific particle electrical or material properties for other methods, can be
generalized to encompass particles of almost any size, shape, density, or compressibility
by generating standing surface acoustic waves within the microfluid channel [2].
Microfluidic mixing [3], which had once been a problem due to laminar flow and
relatively slow diffusion rates, has many solutions within the field of acoustics that utilize
surface acoustic waves to quickly and efficiently mix two fluids into a uniform solution
[4]. Propagating fluids within micro-channels once required bulky syringes and/or
control equipment to ensure proper flow rate, but can now be done with small
piezoelectric transducers [5]. With the arrival of acoustic microfluidics, scientists are
deriving new ways to push the industry forward and create next generation solutions to
solve key problems within the field of microfluidics.

II. LITERATURE REVIEW

1.0 Acoustic Particle Manipulation

Particle separation is often necessary for analyzing chemicals and cells that cannot easily
be separated using standard mechanical processes. Using acoustic waves is an excellent
way to separate, entrap, and direct particles to move in a desired direction, and can be
done with particles of almost any shape, size, or material type. Other popular methods of
particle separation require specific material properties, such as dielectrophoresis (DEP),
which uses an applied electric field to separate particles based on their electrical
properties [6]. Standing surface acoustic wave (SSAW) particle separation works
extremely quickly and efficiently. It also does not require any specific material property
for the process to work. Non surface, bulk acoustic waves can also be used for particle
separation, but it requires larger transducers and requires that the microfluidic channel
material is capable of reflecting acoustic waves between the channel walls [2].
Particle manipulation in microfluidic channels allows for both the separation of
impurities from a fluid and the exchange of a fluid medium. By utilizing different
particles with positive and negative φ values, otherwise known as the acoustic contrast
factor, engineers can create microfluidic devices that separate particles with different
density, compressibility, and size. In the biomedical field, one can harness this
technology to separate various biological particulates such as lipids and red blood cells
[7]. Figure 1.1 shows an example acoustic contrast factor particle separation.

Another way the biomedical field


can harness this technology is by
separating the particles in a
solution from their carrier
medium, such as red blood cells
from plasma. This is especially
important for cases where the
original carrier medium is
Figure 1.1: [7] contaminated. This process can
Particles with +φ values will align at the node while particles with
–φ values will align at the anti-node of a pressure wave. easily be achieved by taking
advantage of laminar flow
properties and utilizing standing acoustic waves. Contaminated fluid can be pumped into
a micro-channel from side channels, and clean carrier fluid can be pumped into the
micro-channel directly from the central channel. When a surface acoustic wave is
propagated through the medium in a way that allows a node to form at the center of the
micro-channel, particles with positive φ values drift into the central part of the flow.
Figure 1.2 displays the difference between particle flow with and without the influence of
a standing acoustic wave [7].

An issue that arises during this process comes


from diffusion between the two fluid
boundaries. With time exposure, a small
amount of the contaminated medium inevitably
diffuses into the clean medium. To minimize
contact time and diffusion rates, flow rate can
be increased. Furthermore, one can increase the
flow rate of the central channel relative to the
Figure 1.2: [7] flow of the side channels to create a diffusion
When the acoustic signal is turned on, particles buffer between the two boundaries of the central
drift into the central fluid due to the acoustic flow [7].
forces pushing particles toward the standing
wave node

Petersson et al. in 2005 measured efficiencies up to 98% in tests separating red blood
cells from contaminated plasma to clean solution. They also found, similar to other
experiments, that acoustic forces were harmless to the cells. If doctors are able to harness
this microfluidic technology, blood that is currently disposed of can be recycled to
patients [7].
Standing surface acoustic waves (SSAWs) can be used to separate and align particles in a
microfluidic channel based on various geometric and material properties such as volume,
density, and compressibility. A SSAW can be created by using two inter-digital
transducers (IDT) on opposite sides of a microfluidic channel. The IDTs are placed on a
piezoelectric substrate to convert an alternating electric field to surface acoustic waves.
When the acoustic pressure frequencies from each transducer are aligned, a standing
wave can be generated and used to capture particles at a node or anti-node of the standing
pressure wave [2].

The same approach used to align particles in a single one dimensional array can be
translated to align particles in multiple one dimensional arrays by using multiple nodes to
trap a series of particles. The line separation can be modeled as a function of frequency,
whereas, higher frequency standing waves cause the distance between parallel particle
lines to decrease [8].

One important factor for determining how fast


particles align to the node of a SSAW is the size of
the particles. Larger particles tend to move toward
the node more quickly. This happens due to the fact
that acoustic wave forces are proportional to
volume, while the viscous forces from the fluid are
proportional to the radius of the particle. This
proportion in force size allows for an interesting
Figure 1.3: [2] solution to the problem of sorting different sizes of
The time required for particles to align to particles into different channels. Engineers can take
the acoustic wave nodes vs. diameter. advantage of the particle alignment speed to quickly
align larger particles into a central channel while
smaller particles, which were not able to align
quickly enough, diverge into branching channels on
the sides of the microfluidic channel [2].

Figure 1.3 shows a graph of the average amount of


time required for various particle sizes to flow into
the SSAW node, and figure 1.4 shows a schematic
Figure 1.4: [2] of small particles diverging into branch channels in
When the SSAW is produced, larger
particles align to the node faster than
the microfluidic device while large particles deviate
smaller particles, allowing smaller to the central branch. Using the
particles to diverge into branching data from the graph of time vs. particle diameter,
channels.
one can engineer a microfluidic channel and adjust
flow-rates to specify which sizes of particles they would like to flow into the central
channel. By adjusting the amplitudes and wavelengths of the standing surface acoustic
waves, one is capable further tuning the tolerance of particle separation [2].

According Jinjie Shi et al. at the Pennsylvania State University in 2009, separation
efficiency has been noted at 90% for small particles and 80% for large particles. This is
an exceptionally good efficiency for particle separation, and it makes this type of
microfluidic particle separation extraordinarily attractive to chemists and biologists who
desire to separate one type of particle or cell from another in a quick and easy way [2].

Surface acoustic waves can also be used to directly guide particles to a certain direction
rather than simply aligning them to a node. Momentum from the acoustic waves can be
transferred to the particles in solution, allowing them to move in the direction of wave
propagation. Microfluidic channels can be designed so that, at a junction, one branch has
a higher flow resistance than the other branch. Similar to electrons in a circuit, particles
in a microfluid desire to go along the path of least resistance. All of the particles coming
toward a junction designed in this way will flow along the path of least resistance unless
an external force is applied which causes them to change paths. With an externally
generated surface acoustic wave created by an IDT, the particles can be pushed toward
the branch of naturally higher resistance and flow along that route [9].

Figure 1.5 shows a diagram of an IDT creating a


surface acoustic wave to guide particles into a
branch of naturally higher flow resistance. Using
this technology, coupled with electrical sensors, has
the potential to create some interesting design
solutions regarding the sorting of particles.
Sensors, such as light sensors that can detect color
and luminescence, can cause the IDT to switch on
or off depending on the particle properties. In this
way, particles can be sensed and easily separated
Figure 1.5: [9] into different channels using directed droplet flow
The difference between natural particle
flow with the IDT turned off and SAW
[9].
induced directed droplet flow with the IDT
turned on.

Further advancements in utilizing two dimensional SSAWs has lead to advancements in a


technology coined as “acoustic tweezers” to create two dimensional arrays and patterns
of particles within a solution using parallel or perpendicular SSAWs. Acoustic tweezers
can use up to 5 orders of magnitude less power than optical tweezers, which utilize a
focused laser beam to precisely move particles. There are many other techniques for
particle manipulation such as optoelectronic tweezers [10], electrophoresis (electrical
field based tweezers) [6], and magnetic tweezers [11], but all of these have some
disadvantages when compared to acoustic tweezers regarding either the scope of particles
that can be aligned, the speed at which they can be aligned, and the power consumption
required [12].

SSAW based tweezers are low power, easy to miniaturize, fast, and effective. Parallel
inter-digital transducers are capable of creating parallel lines of particles along a channel,
whereby the parallel lines of particles settle on the acoustic wave node or anti-node lines.
Likewise, perpendicular inter-digital transducers are capable of forming a two
dimensional pattern of particles, whereby the particles align on an acoustic wave node or
anti-node point within the solution. Figure 1.6 shows an example of acoustic tweezers set
up in a parallel and perpendicular fashion. Once particles are immersed in solution and
the IDT’s are turned on, particles quickly align to the nodes, and the user can alter the
size of the particle groupings just by adjusting the power of the acoustic waves. Higher
power causes the amplitude of the pressure waves to be more extreme, thus, causing the
particles to form more closely together within the node or anti-node [12].

If cells are used as the particles for the 2D patterning,


one worry stems from the acoustic forces damaging the
cells. Research has shown that this isn’t the case. Once
cells reach their resting position in the acoustic standing
wave, they experience almost zero net force acting on
them. The effects of acoustic waves on various sizes of
particles can be calculated by accounting for the
possible forces acting upon them. Particles generally
feel forces from four different sources: acoustic
radiation, viscosity, buoyancy, and gravity. Viscous and
acoustic forces account for the rate of particle grouping,
while gravitational and buoyant effects on particles are
generally negligible due to their relatively equal and
opposite magnitude. Particles with a diameter greater
than 1 micrometer experience acoustic forces much
Figure 1.6: [12] more strongly than viscous forces. Furthermore, as
(a) Parallel (line) tweezers applied acoustic wavelength decreases, acoustic forces
(b) Perpendicular (point) tweezers
are able to overtake viscous forces [12].

The high precision, low power consumption, high tunability, speed, minimalism, and
gentle mechanical nature make acoustic tweezers a useful and innovative tool for many
industries within chemistry and biology [12]. In fact, a number of biological applications
have already been tested in the laboratory to see the effects that acoustic waves have on
living cells. Studies have shown that acoustic trapping does virtually no harm to the
biological cells [13]. The low stress environment of acoustic waves can allow live cells
to live and reproduce freely while trapped in solution, and it allows the user to have
precise control over his or her microenvironment [14].

Three experiments in particular were done at Uppsala University in Sweden by Michael


Evander in 2007 to see how acoustic wave trapping would affect rat spleen cells, yeast
cells, and neural stem. In the first experiment, rat spleen cells were dispersed through a
fluid, and the standing acoustic wave was emitted to see if the cells could be captured.
This experiment performed flawlessly, and the rat spleen cells were trapped almost
instantly. In the second experiment, yeast cells were trapped in solution for six hours to
see how the cells would react to the environment. The yeast cells grew and reproduced
just as expected, and the acoustic wave environment had no visible negative short or long
term effects on the observed cells. In addition, it was concluded from this experiment
that acoustic trapping applied to cells in different mediums of solution would be a
phenomenal way to study growth rates and activity in a highly controlled
microenvironment for further studies. The third experiment, done on neural stem cells,
allowed a fluorescent dye known as “acridine orange” to be filtered into the channel
while they were acoustically trapped. If the cells were still alive and well, they would be
able to absorb the fluorescent dye. After 15 minutes, no perceptible harm was done to the
cells and they remained alive in solution. Moreover, it was concluded that the thermal
environment of an acoustic wave trap is mild enough to create a sustainable living
environment for cells for an indefinite period of time [14].

As well to these three experiments, at Penn State University in 2009, JinJie Shi et al.
performed an experiment using a trapped, two dimensional arrays of e-coli cells for 12
hours. A control sample was heated to 70 degrees Celsius to compare cells which had
obvious visible membrane damage to undamaged cells trapped in the acoustic tweezers.
Results showed no visible signs of membrane damage for the acoustically trapped e-coli
cells, further verifying that using acoustic trapping methods on biological structures
leaves them virtually unharmed [12].

In addition to research based bio-applications of acoustic microfluidics, forensic


applications of acoustic microfluidics show great potential to help solve criminal
investigations. For certain criminal cases, it is necessary for forensic investigators to
separate genetic material and DNA from cells. The current process to analyze samples is
through using a method called differential extraction. The method of differential
extraction is notoriously time consuming, inefficient, and difficult. Utilizing microfluidic
technology and acoustic wave trapping with differential extraction has lead to new
developments in a combined form of analysis called acoustic differential extraction.
Acoustic differential extraction can utilize a large range of sample sizes, ranging from
microliters to milliliters and is capable of separating DNA very quickly. The difference
in size between cells and DNA allows scientists to tune the acoustic waves to hold the
sperm cells in place while allowing DNA molecules to flow freely without becoming
trapped. Using acoustic differential extraction, the time required to analyze a same was
cut down to only 14 minutes, an amount of time that could completely revolutionize
differential extraction methods in forensics and allow microfluidics and lab-on-a-chip
applications to aid criminal investigations [15].

2.0 Acoustic Mixing

Microfluidic mixing is one of the most important processes required to create true lab on
a chip devices. In a microfluidic device, turbulent fluid mixing does not naturally occur
due to highly laminar flow. Diffusion does occur between fluids, but is too kinetically
slow to be very useful for most mixing purposes [16]. A variety of IDT setups can be
used to generate surface acoustic waves to agitate fluid streams and mix them much faster
than diffusion can occur. Single or multiple transducers at the same power output yield
excellent mixing results [4] [16]. Transducers can also be used at varying power outputs,
taking advantage of wave interference, to create a mixing effect known as chaotic
advection [17].
Transducers can be used to generate acoustic waves
capable of thoroughly mixing a series of laminar fluid
streams. By placing a transducer underneath fluid
flow, the generated acoustic waves have enough
energy to disturb the laminar flow and create violent
uniform mixing across the channel. However, if fluid
flow rate is too high, the transducer may not be able
to mix the fluid well enough to create a homogeneous
Figure 2.1: [16]
solution. To alleviate this problem, grid-like Multiple variations of transducer
structures of transducers can be placed underneath the arrays are shown. These transducers
channel to more thoroughly mix multiple streams of are placed below the microfluidic
channel to create acoustic waves
laminar fluids. Figure 2.1 depicts various transducer which mix the fluid above.
patterns that can be used to mix fluid [16].

Surface acoustic waves can be produced to agitate and mix fluids by using an IDT to
excite a piezoelectric substrate. Previous discussions of IDTs in this paper were used to
create standing waves to align particles, but they can also be used to agitate fluids with
proper orientation and power output. SAWs applied to fluid droplets create an effect
called acoustic streaming. In acoustic streaming, the acoustic pressure wave can push
droplets along a surface in the direction of wave propagation [18]. Figure 2.2 depicts
acoustic streaming for a droplet. Internal streaming within a droplet can mix the fluid as

the droplet moves along the surface. If a high


frequency SAW is generated by an IDT placed
directly underneath multiple fluid streams,
perpendicularly to the flow direction, can
efficiently agitate and mix the fluid streams [4-5]
[16]. In addition, SAWs oriented parallel to flow
Figure 2.2: [4]
direction can aid in pumping the fluid along the
Droplet pushed by acoustic streaming channel, similar to the acoustic streaming effect
that can be seen with droplets [19-20].

Chaotic advection [21] can be used to agitate and mix multiple fluids together to allow
mixing of fluids with incredibly low Reynolds numbers [22], less than .1, and high Peclet
numbers [23], greater than 4x105. The Peclet number is the product of the Reynolds
number and the Prandtl number [24]. It is defined as the heat transport by convection
divided by the heat transport by conduction. The Prandtl number is a number describing
the convection of a fluid, and is defined as the quotient between the kinematic viscosity
and the thermal diffusivity. Chaotic advection can be induced using a surface acoustic
wave to create streaming patterns. When two IDTs are operate at separate, variable, and
out of phase powers, they can be used to effectively mix two laminar flowing fluids.
Figure 2.3 shows an experiment using SAWs to
produce chaotic advection mixing. The process
used in these photographs was done using two
tapered IDTs to create interfering SAWs. The
power of one IDT was held constant, while the
power of the other IDT was modulated over various
frequencies to create acoustic pressure disturbance
within the fluid, causing mixing to occur. Each
picture represents a different frequency of power
Figure 2.3: [17]
Two tapered IDTs were used to create mixing oscillation for the second IDT. Power modulation
between fluids. One IDT was held at constant at a frequency of .17Hz is the only frequency that
power, while another IDT had an oscillating
power at 0Hz, .042Hz, .083Hz, .17Hz, .34Hz, could produce a uniform distribution in the
and .68Hz for pictures a, b, c, d, e, and f. experiment [17].

Sectored Fresnel ring transducers allow acoustic waves to be


focused toward the tip of the transducer section. Although they
are generally used to pump or eject fluids, they can also be
arranged in circular or crossing patterns to create an acoustic
mixing effect. Further applications of Fresnel ring transducers
will be discussed in detail later in the paper. Figure 2.4 shows a
Figure 2.4: [25]
Sectored Fresnel rings circular pattern of sectored Fresnel ring mixers [25].
focus acoustic waves to
swirl and mix fluid.

3.0 Acoustic Fluid Pumping / Ejection

Microfluidic systems are often difficult to operate without proper control mechanisms,
which are often large relative to the small microfluidic system. Improving control
mechanisms to change flow rate and pressure is essential to the evolution of microfluidics
because reliance on externally applied pressures from relatively bulky equipment resists
developments of completely integrated lab-on-a-chip devices [26]. New and innovative
ways of propagating fluid through a microfluidic channel using acoustic waves will help
promote microfluidics in mainstream applications and allow engineers to develop self-
contained systems.

In recent research done at Harvard University, Langelier et al. devised a way to control
pressure of a microfluidic system using acoustic wave signals. Their device can control
pressures between 0 and 200 Pascal’s, have a control sensitivity of 10 Pascal’s, and can
potentially reduce the need for external pressure control systems. Figures 3.1 and 3.2
show a drawing of these different sized acoustic resonator cavities and a possible series
of musical notes can be played to induce resonant frequencies in the resonance cavities to
cause fluid motion in the device [26].
The operation of the device is quite unique. Musical
notes cause fluidic motion using resonance cavities to
convert the oscillating pressure into a net directional
flow. These resonance cavities react to acoustic wave
matching their individual resonant frequencies.
Shorter cavities have higher resonant frequencies,
while longer cavities have lower resonant frequencies.
Resonance within the cavities occurs due to the
Figure 3.1: [26] creation of standing waves that arise when the
Resonance cavities resonate at different
frequencies depending on their length. acoustic reflections inside the cavity create perfectly
constructive interference. This constructive
interference occurs when the length of
the cavity is ¼ the wavelength of the
acoustic signal and on every
consecutive odd integer harmonic.
Groups of these acoustic resonators
can be clustered together and used as
multiple pumps for a microfluidic
system. By generating a code of
Figure 3.2: [26] musical notes, information can be sent
Signals generated from musical notes can be played into the from a speaker to the resonators and
resonance cavities to cause fluid motion within the
microfluidic channels.
cause the fluids to move according to
the notes played by the speaker [26].

Rectifiers are structures that cause the oscillating resonance to impart a unidirectional
fluid motion instead of allowing the fluid to oscillate back and forth with the acoustic
wave. It works by using a flap valve which only opens during one direction of fluid input
and closes if the fluid tries to move in the opposite direction. The flap valve is
engineered so that oscillating pressure is strong enough to open the flap valve only if
resonance is achieved within the cavity. Using a computer to control the frequency,
duration, and amplitude of the acoustic signal allows for precise control over the velocity
and distance traveled by the fluid [26].

It is possible to further scale down acoustic fluid propagation by using piezoelectric


transducers of the vibrations from a membrane. If this small scale system can be
developed, lab on a chip devices can be simplified and have more widespread
applications due to the ease of control and lack of large external equipment [26].

Modern day ink-jet printers use nozzle based printing techniques in which the ink
ejection direction is fixed perpendicular to the nozzle surface. This scheme has draw
backs due to the fact that the printing head must be moved if the user desires to use more
than one type of ink. Utilizing focused acoustic waves with a nozzleless ejector design
allows atomized ink droplets to be precisely ejected in any desired direction. Without the
acoustic waves to guide the liquid, the nozzleless system will spray ink in random
directions. In addition, using this technique allows for multiple inks to be used without
requiring the printer head to move [27].
Both acoustic lenses and constructive interference can be used to focus acoustic waves.
Constructive interference, also known as self-focusing, can be produced by using fresnel
ring electrodes on a piezoelectric surface to generate acoustic pressure orthogonal or
angled to the surface of the substrate. Piezoelectric substances, along with the sectioned
annular electrodes, allow the nozzleless printer to direct the ink droplet. Acoustic waves
generated by the piezoelectric material create a constructive pressure interference wave
orthogonal to the ink surface and allow a small particle of ink to be ejected from the
liquid due to the accumulated pressure [27].

As shown in figure 3.3, due


to the ring design, ink
droplets can be ejected at
an angle from the surface of
the ejector. Since the
electrodes create larger
acoustic waves than the pie
slice that has no electrodes,
Figure 3.3: [26] the droplet direction ejects
Openings in the electrode rings on top of the piezoelectric substrate cause toward the pie slice shape
an non-symmetric distribution of acoustic waves, causing liquid particles to [27]
eject at an angle from the surface.

Changing the angle of the open slice allows for ink droplets to be ejected at many
different angles. The smaller the slice angle, the closer the ejection angle comes to 90
degrees perpendicular to the ink surface. For ZnO piezoelectric material, as the slice
angle increases to 90 degrees the ink ejection angle can change all the way to 62.5
degrees from the surface. Using this process, ink drops of 10 micrometers or larger in
diameter can be ejected using a nozzleless, piezoelectric, Fresnel ring electrode
stimulated device [27].

It is also possible to create a nozzleless device with multiple ejection angles based on a
single concentric annular ring design. One can change the acoustic vibrations by
blocking one of the electrode slices. With an electrode design using 8 slices, 8 different
ejection directions can be used for the ink. By turning off a slice of an electrode, the ink
drop becomes influenced to eject in the direction of the non-powered electrode due to a
lack of acoustic pressure in that area. Figure 3.4 shows a simple diagram explaining this
process of ink ejection in the direction of the un-powered electrode ring slice [27].

Using these sectioned electrodes, multi-


directional ink ejection has been
achieved. If a large array of these
piezoelectric ink ejectors was made, a
printer could be designed to print with
ultra high quality and resolution.

Figure 3.4: [26]


Fresnel rings are sectioned to allow ink droplets to eject
at 8 different angles surrounding the surface of the
substrate simply by adjusting which section is turned off.
The ejection rate of the ink droplets is governed by the radio frequency pulse repetition,
and ink droplets of sizes ranging from a few micrometers to tens of micrometers have
been ejected with radio pulses ranging from a few microseconds to a few hundred
microseconds. Moreover, liquid temperature during the acoustic ejection process remains
constant, that is, the acoustic waves do not affect the temperature of the ejected liquid.
This benefit makes self-focusing acoustic transducer ejectors a good tool for handling
temperature sensitive liquids which need to be printed [26]. Acoustic based liquid jetting
has already found commercial applications to collect precisely measured nano-liter fluid
volumes [28].

As previously discussed, SAW streaming effects applied to droplets can push them along
a surface. When SAWs are directed along the flow of fluid in a microchannel, they are
capable of pushing entire streams of fluid. At higher powers, SAW induced fluid
propagation can create jetting [5] [18] or even atomization of liquid [5] [29-30].
Moreover, SAWs can push liquid through microchannels at an extraordinarily fast rate,
up to speeds ranging from1cm/s and 10cm/s [5].

In addition to using Fresnel rings to eject [27] and


mix [25] fluids, they can also create an acoustic
streaming effect used to propagate fluid through the
channel. Each Fresnel ring works by producing a
Figure 3.5: [25]
Sectioned Fresnel rings can pump fluid in a focused surface acoustic wave which propagates in
microchannel by focusing acoustic waves along the direction moving from the largest ring radius to
the flow direction.
the smallest ring radius. An array of sectioned
electrode Fresnel rings on a piezoelectric substrate is capable of producing powerful
acoustic waves when patterned to create constructive interference. Figure 3.5 shows an
example of a sectioned Fresnel ring array that can be used to push liquid through a
microchannel [25].

Surface acoustic waves are capable of producing


huge stresses and accelerations on fluids when
actuators are used at high powers. As the power
output approaches 1 watt, they SAWs are capable of
accelerating fluid up to 107 to 108 m2/s, so fast that
the liquid can be converted into atomized drops
between 1 and 10 micrometers in diameter [5] [29-
30]. Using this technique to create nano-sized
particles can be widely used throughout a variety of
industries such as printing, drug delivery, agricultural
Figure 3.6: [5] spraying, and fuel injection. SAWs are also an
Polymeric nanoparticles generated from excellent way to generate polymeric nano-particles of
surface acoustic wave fluid atomization 150 to 200 nanometers as shown in figure 3.6 [5].
III. CONCLUSIONS

Acoustic microfluidic technology has steadily improved as new and creative applications
have been developed and researched. Acoustics has proven itself to be one of the most
useful ways of separating particles, mixing fluids, and propagating microfluidic solutions,
in addition to its capabilities of miniaturizing microfluidic control systems. Biology,
biomedicine, chemistry, forensics, ink-jet printing, and nano-particle fabrication are just a
few of the many potential avenues that acoustic microfluidics can assist. Ultimately,
acoustics is just one solution of many to solve problems within the field of microfluidics,
but few other methods provide the generality, tunability, and development potential of
utilizing acoustic waves.

IV. REFERENCES

[1] Gualtieri, J.G.; Kosinski, J.A.; Ballato, A., (1994). Piezoelectric materials for acoustic wave applications.IEEE
Transactions on Ultrasonics, Ferroelectrics and Frequency Control, 41(1), pp.53-59, Jan 1994

[2] Shi, J., Huang, H., Stratton, Z., Huang, Y., & Huang, T. J. (2009). Continuous particle separation in a microfluidic
channel via standing surface acoustic waves (SSAW) . Lab on a chip, 9, pp.3354-3359.

[3] Nam-Trung Nguyen and Zhigang Wu (2005). Micromixers – a review. Journal of Micromechanics &
Microengineering. 15 R1-R16 doi: 10.1088/0960-1317/15/2/R01

[4] Sritharan, K., Strobl, C., Schneider, M., Wixforth, A., & Guttenberg, Z. (2006). Acoustic mixing at low Reynold's
numbers. Applied Physics Letters, 88(054102), pp.1-3.

[5] Yeo, L., & Friend, J. (2009). Ultrafast microfluidics using surface acoustic waves. Biomicrofluidics, 3(012002),
pp.1-23.

[6] Gascoyne P. R. C. and Vykoukal J., (2002). Electrophoresis 2-1. 23(1973) doi: 10.1002/1522-
2683(200207)23:13<1973::AID-ELPS1973>3.0.CO;

[7] Petersson, F., Nilsson, A., Jonsson, H., & Laurell, T. (2005). Carrier Medium Exchange through Ultrasonic Particle
Switching in Microfluidic Channels. Analytical Chemistry, 77(5), pp.1216-1221.

[8] Wood, C., Evans, S., Cunningham, J., O'Rorke, R., Walti, C., & Davies, A. (2008). Alignment of particles in
microfluidic systems using standing surface acoustic waves. Applied Physics Letters, 92(044104), pp.1-3.

[9] Thomas Franke, Adam R. Abate, David A. Weitz and Achim Wixforth. (2009) Surface acoustic wave (SAW)
directed droplet flow in microfluidics for PDMS devices. Lab on a Chip, 9, pp. 2625-2627

[10] Ohta, Aaron T., Chiou, Pei-Yu, Han, Tae H., Liao, James C., Bhardwaj, Urvashi, McCabe, Edward R.B., et al.
(2007). Dynamic cell and microparticle control via optoelectronic tweezers. UC Berkeley: Retrieved from:
http://escholarship.org/uc/item/2hc7425b

[11] Charlie Gosse and Vincent Croquette. (2002) Magnetic Tweezers: Micromanipulation and Force Measurement at
the Molecular Level. Biophysical Journal – 1, 82(6), pp. 3314-3329

[12] Shi, J., Ahmed, D., Mao, X., Lin, S. S., Lawit, A., & Huang, T. J. (2009). Acoustic tweezers: patterning cells and
microparticles using standing surface acoustic waves (SSAW). Lab on a chip, 9, pp.2890-2895.

[13] Castillo, J., Dimaki, M., & Svendsen, W. E. (2008). Manipulation of biological samples using micro and nano
techniques . Integrative Biology, 1, pp.30-42.

[14] Evander, M., Johansson, L., Lilliehorn, T., Piskur, J., Lindvall, M., Johansson, S., et al (2007). Noninvasive
Acoustic Cell Trapping in a Microfluidic Perfusion System for Online Bioassays. Analytical Chemistry,
79(7), pp.2984-2991.

[15] Norris, J. V., Evander, M., Horsman-Hall, K. M., Nilsson, J., Laurell, T., & Landers, J. P. (2009). Acoustic
Differential Extraction for Forensic Analysis of Sexual Assault Evidence. Analytical Chemistry, 81(15),
pp.6089-6095.

[16] Yaralioglu, G. G., Wygant, I. O., Marentis, T. C., & Khuri-Yakub, B. T. (2004). Ultrasonic Mixing in Microfluidic
Channels Using Integrated Transducers. Analytical Chemistry, 76, pp.3694-3698.

[17] Frommelt, T., Kostur, M., Wenzel-Schafer, M., Talkner, P., Hanggi, P., & Wixforth, A. (2007). Microfluidic
Mixing via Acoustically Driven Chaotic Advection. Physical Review Letters, 100(034502), pp.1-4.

[18] Beyssen, D.; Brizouala, L. L.; Elmazriaa, O. & Alnota, P., (2006). Microfluidic device based on surface acoustic
wave . Sensors and Actuators B: Chemical, 118(1-2), pp.380-385.

[19] Renaudin, A., Tabourier, P., Zhang, V., Camart, J., & Druon, C. (2006). SAW nanopump for handling droplets in
view of biological applications . Sensors and Actuators B: Chemical, 113(1), pp.389-397.

[20] Guttenberg, Z., Rathgeber, A., Keller, S., Radler, J., Wixforth, A., Kostur, M., et al (2004). Flow profiling of a
surface-acoustic-wave nanopump. Physics Review Letters, 70(056311), pp.1-10.

[21] Advection: transport of something from one region to another. (n.d.). WW2010 (the weather world 2010 project):.
Retrieved November 15, 2009, from http://ww2010.atmos.uiuc.edu/(Gh)/guides/mtr/af/adv/adv.rxml

[22] Calculating the Reynolds Number. (n.d.). Rice. Retrieved November 15, 2009, from
http://www.owlnet.rice.edu/~chbe402/proj05/jigarb/Reynolds_number.htm

[23] Calculating the Peclet Number. (n.d.). Rice. Retrieved November 15, 2009, from
www.owlnet.rice.edu/~chbe402/proj05/jigarb/peclet_number.htm

[24] Weisstein, E. W. (n.d.). Prandtl Number -- from Eric Weisstein's World of Physics. ScienceWorld. Retrieved
November 15, 2009, from http://scienceworld.wolfram.com/physics/PrandtlNumber.html

[25] Yu, H., Kwon, J. W., & Kim, E. S. (2006). Microfluidic mixer and transporter based on PZT self-focusing acoustic
transducers. Journal of Microelectromechanical Systems, 15(4), pp.1015-1024.

[26] Langelier, S. M., Chang, D. S., Zeitoun, R. I., & Burns, M. A. (2009). Acoustically driven programmable liquid
motion using resonance cavities. PNAS, 106, 12617-12622; doi:10.1073/pnas.0900043106.

[27] Kwon, J. W., Yu, H., Zou, Q., & Kim, E. S. (2006). Directional droplet ejection by nozzleless acoustic ejectors
built on ZnO and PZT. Journal of Micromechanics and Microengineering, 16, pp.2697-2704.

[28] Olechno, J. (2007, December 13). Acoustic Droplet Ejection. SciVee | Making Science Visible. Retrieved
November 15, 2009, from http://www.scivee.tv/node/4522

[29] Chono, K., Shimizu, N., Matsui, Y., Kondoh, J., & Shiokawa, S. (2004). Development of Novel Atomization
System Based on SAW Streaming. Japanese Journal of Applied Physics, 43(5B), pp.2987-2991.

[30] Kurosawaa, M., Watanabea, T., Futamia, A., & Higuchia, T. (1995). Surface acoustic wave atomizer . Sensors and
Actuators A: Physical, 50(1-2), pp.69-74.

You might also like