You are on page 1of 11

Modeling and Simulation of a Wireless Passive

Thermopneumatic Micromixer

Marwan Nafea, Nasarudin Ahmad, Ahmad Ridhwan Wahap,


and Mohamed Sultan Mohamed Ali(&)

Faculty of Electrical Engineering, Universiti Teknologi Malaysia,


81310 Skudai, Johor, Malaysia
nmmarwan3@live.utm.my,
{e-nasar,aridhwanwahap}@utm.my, sultan_ali@fke.utm.my

Abstract. This paper presents modeling and simulation of a wirelessly-


controlled thermopneumatic zigzag micromixer. The micromixer is operated by
selectively activating two passive wireless heaters with different resonant fre-
quencies using an external magnetic field. Each heater is responsible for heating
an air-heating chamber that is connected to a loading reservoir through a
microdiffuser element, while the solutions pumped from each reservoir are
mixed in a zigzag micromixing element that ends with an outlet hole. The
performance of the micromixer is analyzed using finite element method, and
mixing is investigated over a low range of Reynold’s number (Re) ⩽ 10 that is
suitable various biomedical applications. The optimal activation switching time
of the heaters is 10 s, at which the micromixer achieves a maximum mixing
efficiency of *96.1%, after *65 s. The micromixer provides mixing-ratio
controllability with a maximum flow rate and pressure drop of *3.4 µL/min
and *385.22 Pa, respectively.

Keywords: Finite element method  Microfluidic  Micromixer 


Thermopneumatic  Wireless actuation

1 Introduction

Micromixers play a core role in various biomedical applications, including point-of-


care [1], lab-on-a-chip [2], and drug delivery systems [3]. These systems require
precise transportation, mixing, and separation of miniature volumes of fluids within
small (typically sub-millimeter)-scale geometries, which makes traditional methods of
mixing fluids inapplicable. In addition, fluids flow is typically laminar in these systems,
causing a low Reynold’s number (Re) that consequently poses a challenge for sample
mixing [4]. At low Re values, mixing is governed by molecular diffusion, which
increases the length of the mixing channel and the time required to achieve a homo-
geneously mixed solution [5].
Micromixers are generally classified as active or passive, according to their mixing
mechanisms [6]. Active micromixers require external forces to perform the mixing
process, such as acoustics, pressure, temperature, electroosmosis, magnetohydrody-
namics, electrohydrodynamics, and dielectrophoretics [5]. On the other hand, passive

© Springer Nature Singapore Pte Ltd. 2017


M.S. Mohamed Ali et al. (Eds.): AsiaSim 2017, Part I, CCIS 751, pp. 312–322, 2017.
DOI: 10.1007/978-981-10-6463-0_27
Modeling and Simulation of a Wireless Passive Thermopneumatic Micromixer 313

micromixers do not require external energy to mix the fluids. This makes passive
micromixers more favorable in microfluidic systems due to their simple fabrication and
implementation as compared to active micromixers [7]. Passive micromixers depend on
molecular diffusion or chaotic advection principles, which are realized through the use
of injection, droplet, parallel and serial lamination, and chaotic advection micromixers
[7]. Chaotic advection micromixers are reported to perform efficiently over a wide
range of Re values when compared to other types of passive micromixers [7]. Chaotic
advection can be generated by modifying the shape of the microchannel in such a way
that the laminar flow continuously splits, folds, stretches, and breaks over the
cross-section of the microchannel. This can be realized using zigzag microchannels,
intersecting microchannels, twisted microchannels, convergent–divergent microchan-
nels, three-dimensional structures, staggered herringbone structures, Tesla structures, or
by inserting grooves, obstacles, or embedded barriers in the microchannel walls [5].
Zigzag microchannel type has been widely investigated due to its high mixing
index and ease of fabrication. Numerical and experimental studies of mixing in planar
zigzag microchannels were performed by Mengeaud et al. [8]. The results indicated a
transition in the mixing performance at a critical Re of *10. Increasing Re led to form
vortices at the corners of the microchannel, which therefore improved the mixing
performance and reduced the diffusion length. In a recent work, a numerical study
was performed on a micromixer that is composed of trapezoidal channels in a zigzag
and split–recombine arrangement [9]. The optimal geometry had a mixing
efficiency >85% for Re values in the range of 0.1–80, and reached up to 90% when Re
⩽ 0.9 and Re ⩾ 20.
Time-dependent pulsatile flows have been reported as a potential approach to
achieve fast and efficient mixing. This approach was investigated by Glasgow et al.
[10] using a low-frequency sinusoidal flow superimposed on a steady flow. The results
indicated that the mixing performance with the pulsed sinusoidal flow was greatly
improved as compared to the flow without pulsing.
However, these devices are limited to the use of on-board power supplies [1] or
bulky magnetic cores [11], which limits the ability to implement these devices in
portable or implantable systems. Passively actuated actuators can be utilized to address
these issues while offering a smaller size, lower cost, higher robustness, and higher
biocompatibility as compared to actively actuated actuators [12]. Wireless radiofre-
quency (RF) magnetic field actuation method has been utilized in several passively
actuated microfluidic applications, such as multiple shape-memory-alloy actuators
control of a microsyringe device [13], implantable drug delivery devices [14], and a
thermopneumatic micropump for biomedical applications [15].
This paper presents an extended modeling and simulation study of a passive zigzag
thermopneumatic micromixer that is wirelessly controlled using an external RF mag-
netic field [16]. Dynamic heat transfer, thermopneumatic actuation and mixing per-
formance are analyzed using finite element method (FEM). The results suggest that the
developed micromixer is able to provide mixing-ratio controllability and can be
potentially implemented in biomedical applications.
314 M. Nafea et al.

2 Design and Working Principle

Figure 1 illustrates a schematic diagram of the developed the developed thermopneu-


matic micromixer. The micromixer consists of two inductor–capacitor (LC) wireless
heaters that have identical planar spiral inductors with dimensions of 5 mm  6 mm,
while the dimensions of each capacitor are 5 mm  2 mm and 5 mm  3 mm,
respectively. Therefore, the heaters have different resonant frequency (fr) values
of *100 and *130 MHz, respectively. Each heater is located beneath two air-heating
chambers that have similar dimensions to the spiral inductors. Each chamber is con-
nected to a 5 mm-diameter loading reservoir through a 10° microdiffuser element to
rectify the flow. The reservoirs are connected through a T-inlet with a width of 0.3 mm
and a mixing length of 2 mm to a zigzag micromixing element that ends with a 1 mm
outlet hole.

Fig. 1. Schematic diagram of the developed micromixer with an external coil (including a
close-up view of the fluid-directing channel)

An AC current is generated in the LC circuit when it is exposed to an external AC


magnetic field, due to the electromotive force induced by the magnetic field. Thus, the
power, P, consumed in the LC circuit is given as [17]:

Rv2
PðxÞ ¼ ð1Þ
½R þ jðxL  1=xCÞ2

where R is the parasitic resistance of the circuit, v is the electromotive force, x is the
angular frequency of the AC current, and L and C are the inductance and capacitance of
the circuit, respectively. The maximum power transferred to the LC circuit when the
pffiffiffiffiffiffi
field frequency is tuned to the resonant frequency of the circuit, fr ¼ 1=2p LC .
Therefore, the Joule heat generated in each LC circuit reaches its maximum value at the
corresponding fr of the LC circuit, and the steady state temperature, TSS, can be
expressed as [17]:
Modeling and Simulation of a Wireless Passive Thermopneumatic Micromixer 315

RT v2 =R
TSS ¼ ð2Þ
1 þ aR RT v2 =R

where RT and aR are the thermal resistance to the surroundings and the temperature
coefficient of the resistance of the circuit, respectively. The generated heat is transferred
by conduction from each LC circuit to the corresponding air-heating chamber through
the polydimethylsiloxane (PDMS) layer. The rate of conductive heat transfer, q, is
obtained using Fourier’s law [18]:

DT
q ¼ kA ð3Þ
Dx
where k is the thermal conductivity of the PDMS, A is the surface contact area between
each LC heater and the corresponding air-heating chamber, and Dx and DT represent
the thickness of the PDMS layer and the temperature difference across it, respectively.
As heat is transferred to the walls of the air-heating chamber, the temperature of the
enclosed air rises, causing the volume of air to expand. This can be generally expressed
by the ideal gas equation while assuming that the heating process is isobaric:

V0
DV ¼ DT ð4Þ
T0

where DV = V1−V0 is the amount of expanding volume, DT = T1−T0 is the elevation


of the chamber temperature, V0 is the volume of the gas at the room temperature, T0
(23 °C in this work), and V1 is the volume at the elevated temperature, T1 [15].
Therefore, the expanded volume of air pushes the liquid from the loading reservoir to
the T-inlet, causing them to mix at the zigzag mixing element.
The volume expansion of the air inside the air-heating chambers causes a defor-
mation in their top walls. This is due to the fact that the high thickness of the side walls
(Fig. 1), and that the bottom walls are attached to the LC heaters that are fabricated out
of a polyimide film (  47 µm). This film is about three orders of magnitude stiffer than
PDMS [19], making it rigid and greatly reducing the deformation of the bottom walls
of the air-heating chambers. Thus, the deformation, s, of each top wall of the chambers
can be approximated as:

3DV 3V0 DT
s  ð5Þ
AW AW T0

where AW is the area of the top wall of the chamber. The deformation is greatly reduced
by increasing the thickness of the top walls of the chambers to 2 mm. Thus, the
majority of the expanded volume of the air is applied on the inlet of the microdiffuser
element, which significantly reduces the pumping loss.
316 M. Nafea et al.

3 Finite Element Analysis

Numerical simulation of a three-dimensional model of the developed micromixer was


carried out using COMSOL Multiphysics®. Heat transfer, solid mechanics, laminar
flow, and transport of diluted species physics were used in the simulation process. It
was assumed that there was no thermal loss to the surroundings, and the initial air
temperature was assumed to be 23 °C.
The temperature of the LC wireless heaters (heat source) was set according to their
experimentally measured temperature [20]. FEM was implemented to solve the flow
and mixing governing equations, which are the Navier-Stokes, continuity, and
convection-diffusion equations, as follows:
 
@u
q þ ðu:rÞu ¼ rp þ lr2 u ð6Þ
@t

ru ¼ 0 ð7Þ

@c
¼ Dr2 c  u:rc ð8Þ
@t

where u is the fluid velocity, D is the diffusion coefficient, and c is the molar con-
centration. The fluid density, dynamic viscosity, and diffusion coefficient were set as
999.8 kg/m3, 10−6 m2/s (the kinematic viscosity of water at room temperature), and
10−10 m2/s, respectively. The boundary conditions were set as follows: molar con-
centrations of 1 mol/m−3 at the inlet of reservoir 1 and 0 mol/m−3 at the inlet of
reservoir 2, no-slip at solid walls, an equal volume flow rate at each microdiffuser inlet,
and zero pressure at the outlet. The flow rate, Q, can be determined as a function of
pressure using the Hagen–Poiseuille relationship:

DpAC r 2
Q¼ ð9Þ
8lL

where AC is the cross-sectional area of the channel, r is the diagonal length of the
channel, and L is the length of the channel. The fluid dynamics in the microscale are
characterized by a dimensionless parameter, Reynolds number (Re) that is defined as
[8]:

quDh
Re ¼ ð10Þ
l

where Dh is the hydraulic diameter of the channel. The standard deviation, r, of the
molar concentration across the channel in a cross-section at any specific longitudinal
location, and the mixing efficiency, η, at the outlet cross-section are calculated as
follows [9]:
Modeling and Simulation of a Wireless Passive Thermopneumatic Micromixer 317

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1X n
r¼ ðci  cÞ2 ð11Þ
n i¼1
sffiffiffiffiffiffiffiffiffi!
r2
g¼ 1  100 ð12Þ
r2max

where ci is the molar concentration at a point i, c is the mean molar concentration of a


fully mixed solution, n is the total number of sample points inside the cross-section,
and rmax is the maximum standard deviation at the inlet of the main channel. A mixing
efficiency of 0% indicates completely separated streams (r = rmax), while a mixing
efficiency of 100% indicates completely mixed streams (r = 0) [20].

4 Results and Discussion

In this work, a time-dependent advanced multi-level multi-discretization GMRES


solver was used to simulate the performance of the developed micromixer. Only heater
1 was activated with a power of 60 mW to demonstrate the heat transfer and thermal
expansion behaviors. The heater was activated for 50 s until its temperature reached
50 °C (experimentally measured) [20] and then deactivated for 230 s until its tem-
perature reached room temperature (23 °C). The temperature distribution profile of the
simulated three-dimensional model of the micromixer after activating heater 1 for 50 s
is illustrated in Fig. 2.

Fig. 2. A three-dimensional model of the micromixer showing the temperature distribution


profile 50 s after activating heater 1.

The generated temperature in heater 1 is transferred by conduction to the bottom


wall of air chamber 1, which therefore increases the temperature of the enclosed air.
This behavior is demonstrated in Fig. 3, which shows that the temperature of the air
chamber follows the change of temperature of the heater and reaches a maximum value
of 42.5 °C after 50 s.
318 M. Nafea et al.

Fig. 3. The simulated temporal response of the temperature of heater 1 and the temperature and
resultant deformation of air chamber 1 after 50 s from activating heater 1.

A deformation on the top surface of the air chamber occurred as the temperature of
the enclosed air was increased, due to the expansion of the enclosed air. A maximum
deformation of *50 µm was generated after 50 s. Then it started to decrease after
deactivating the heater until reaching its initial condition when the temperature of the
air chamber returned to the room temperature.
To investigate the pumping and mixing performances of the micromixer, heaters 1
and 2 were operated using two different activation and deactivation patterns, as illus-
trated in Fig. 4(a). The resultant mixing efficiencies at the outlet when operating the
micromixer for 140 s using these two activation patterns are presented in Fig. 4(b). It
can be indicated that a maximum mixing efficiency of *96.1% was achieved when
using each activation pattern. However, activation pattern 1 generated a more stable
mixing efficiency and a faster time (*65 s) required to reach the maximum mixing
efficiency. On the other hand, activation pattern 2 showed a less stable mixing per-
formance that fluctuated between *80.9% and *96.1% and required a higher
response time (*70 s). It can be concluded that activation pattern generates a better
performance than activation pattern 2 due to the fact that the latter one allows unwanted
amounts of fluids to pass from one reservoir to the mixing channel without allowing a
similar opposing amount from the other reservoir within a short period of time.
The mixing efficiency at the outlet and the pressure drop were examined over a range
of Re ⩽ 10, as shown in Fig. 5. It can be indicated that the highest mixing efficiency
was *96.15% at Re = 0.1, while it was it its lowest value (81.14%) at Re = 2. Beyond
that range, the mixing efficiency was improved up to 90.36% at Re = 10. This variance
in the mixing performance is related to the transverse velocity components that are
increasingly generated around the angle of the zigzag element at higher Re values.
However, at low Re values, the transverse flow is negligible, which means that mixing is
dominated by the molecular diffusion as a result of the low velocity and the high
residence time. Further observations of Fig. 5 show that the pressure drop increased
with Re, with a maximum pressure drop of *385.22 Pa at Re = 10.
Modeling and Simulation of a Wireless Passive Thermopneumatic Micromixer 319

Fig. 4. Effect of activation patterns on the mixing efficiency. (a) Two activation patterns.
(b) Numerical results of the mixing efficiency using the two activation patterns over 140 s.

Fig. 5. Mixing efficiency and pressure drop at the outlet for different Re values.

The distribution of the mixing efficiency along the axial mixing length was
investigated at three different Re values, i.e. 0.1, 1, and 10, as shown in Fig. 6. The
results indicated that the three distributions of the mixing efficiencies follow the same
320 M. Nafea et al.

trend. It can be observed that the mixing efficiencies were greatly improved at each
corner of the zigzag mixing element, i.e. at 500, 800, and 1100 lm of the axial mixing
length. Furthermore, it can be noted that the distribution of the mixing efficiency at
Re = 0.1 was the best along the whole axial mixing length, while it was the lowest at
Re = 1.

Fig. 6. Mixing efficiency along the axial mixing length for Re = 0.1, 1 and 10.

From Figs. 5 and 6, It can be indicated that the mixing efficiency can be improved
beyond this range of Re values. However, the developed micromixer is intended to be
used in typical microfluidic biomedical applications, where high flow velocities can
cause rupture of cells, where the maximum flow rate generated at Re = 10 was *3.4
µL/min. In addition, the operating temperature required to achieve higher Re values is
limited to avoid any potential damage that might occur to the device.

5 Conclusion

A wirelessly-controlled thermopneumatic zigzag micromixer was demonstrated in this


paper. The micromixer was operated using two passive heaters. Each heater generated
heat when an external field frequency was tuned to its corresponding resonant fre-
quency. Dynamic heat transfer, thermopneumatic actuation, and mixing performance
were analytically presented and then verified using FEM as a demonstration of the
proposed mixing technique. The proposed micromixer was able to achieve a maximum
mixing efficiency of *96.1%, with a maximum flow rate and pressure drop of *3.4
µL/min and *385.22 Pa, respectively. Two activation switching patterns were
investigated to find the optimal operating time of the two heaters. It was found that the
optimal switching time is 10 s, at which the micromixer achieved a steady value of the
maximum mixing efficiency after *65 s.
The performance of the micromixer can be further improved by implementing
metamodeling approach to find the optimal device design and activation pattern [21].
Modeling and Simulation of a Wireless Passive Thermopneumatic Micromixer 321

Acknowledgments. This work was supported by Research University Grant (10H40 & 14H31)
from Universiti Teknologi Malaysia; and Fundamental Research Grant Scheme (FRGS) &
Prototype Development Research Grant Scheme (PRGS) from Ministry of Higher Education
Malaysia. M. Nafea acknowledges the financial support from the Malaysian Technical Coop-
eration Programme (MTCP).

References
1. Maleki, T., Fricke, T., Quesenberry, J., Todd, P., Leary, J.F.: Point-of-care, portable
microfluidic blood analyzer system. In: International Society for Optics and Photonics,
Microfluidics, BioMEMS, and Medical Microsystems X, pp. 82510C-82513, California
(2012)
2. Rajabi, N., Bahnemann, J., Tzeng, T.-N., Platas Barradas, O., Zeng, A.-P., Müller, J.:
Lab-on-a-chip for cell perturbation, lysis, and efficient separation of sub-cellular components
in a continuous flow mode. Sens. Actuators, A 215, 136–143 (2014)
3. Belliveau, N.M., Huft, J., Lin, P.J.C., Chen, S., Leung, A.K.K., Leaver, T.J., Wild, A.W.,
Lee, J.B., Taylor, R.J., Tam, Y.K., Hansen, C.L., Cullis, P.R.: Microfluidic synthesis of
highly potent limit-size lipid nanoparticles for in vivo delivery of siRNA. Mol. Ther. -Nucl.
Acids 1, 37 (2012)
4. Hossain, S., Kim, K.-Y.: Mixing analysis of passive micromixer with unbalanced three-split
rhombic sub-channels. Micromachines 5, 913–928 (2014)
5. Lee, C.-Y., Chang, C.-L., Wang, Y.-N., Fu, L.-M.: Microfluidic mixing: a review. Int.
J. Mol. Sci. 12, 3263–3287 (2011)
6. Li, X., Chang, H., Liu, X., Ye, F., Yuan, W.: A 3-D overbridge-shaped micromixer for fast
mixing over a wide range of reynolds numbers. J. Microelectromech. Syst. 24, 1391–1399
(2015)
7. Nam-Trung, N., Zhigang, W.: Micromixers—a review. J. Micromech. Microeng. 15, R1
(2005)
8. Mengeaud, V., Josserand, J., Girault, H.H.: Mixing processes in a zigzag microchannel:
finite element simulations and optical study. Anal. Chem. 74, 4279–4286 (2002)
9. Le Hai, T., Bao Quoc, T., Le Hoa, T., Tao, D., Trung Nguyen, T., Frank, K.: Geometric
effects on mixing performance in a novel passive micromixer with trapezoidal-zigzag
channels. J. Micromech. Microeng. 25, 094004 (2015)
10. Glasgow, I., Lieber, S., Aubry, N.: Parameters influencing pulsed flow mixing in
microchannels. Anal. Chem. 76, 4825–4832 (2004)
11. Zhu, G.-P., Nguyen, N.-T.: Rapid magnetofluidic mixing in a uniform magnetic field. Lab
Chip 12, 4772–4780 (2012)
12. Joung, Y.-H.: Development of implantable medical devices: from an engineering
perspective. Int. Neurourol. J. 17, 98–106 (2013)
13. Mohamed Ali, M.S., Takahata, K.: Wireless microfluidic control with integrated
shape-memory-alloy actuators operated by field frequency modulation. J. Micromech.
Microeng. 21, 075005 (2011)
14. Smith, S., Tang, T.B., Terry, J.G., Stevenson, J.T.M., Flynn, B.W., Reekie, H.M., Murray,
A.F., Gundlach, A.M., Renshaw, D., Dhillon, B., Ohtori, A., Inoue, Y., Walton, A.J.:
Development of a miniaturised drug delivery system with wireless power transfer and
communication. IET Nanobiotechnol. 1, 80–86 (2007)
15. Chee, P.S., Minjal, M.N., Leow, P.L., Mohamed Ali, M.S.: Wireless powered
thermo-pneumatic micropump using frequency-controlled heater. Sens. Actuators A:
Physical 233, 1–8 (2015)
322 M. Nafea et al.

16. Nafea, M., AbuZiater, A., Faris, O., Kazi, S., Mohamed Ali, M.S.: Selective wireless control
of a passive thermopneumatic micromixer. In: 2016 IEEE 29th International Conference on
Micro Electro Mechanical Systems (MEMS), pp. 792–795. IEEE Press, Shanghai, China
(2016)
17. Mohamed Ali, M.S., Takahata, K.: Frequency-controlled wireless shape-memory-alloy
microactuators integrated using an electroplating bonding process. Sens. Actuators A:
Physical 163, 363–372 (2010)
18. Chee, P.S., Nafea, M., Leow, P.L., Mohamed Ali, M.S.: Thermal analysis of wirelessly
powered thermo-pneumatic micropump based on planar LC circuit. J. Mech. Sci. Technol.
30, 2659–2665 (2016)
19. Guo, L., Guvanasen, G.S., Liu, X., Tuthill, C., Nichols, T.R., DeWeerth, S.P.: A
PDMS-based integrated stretchable microelectrode array (isMEA) for neural and muscular
surface interfacing. IEEE Trans. Biomed. Circ. Syst. 7, 1–10 (2013)
20. Nafea, M., AbuZaiter, A., Kazi, S., Mohamed Ali, M.S.: Frequency-Controlled Wireless
Passive Thermopneumatic Micromixer. J. Microelectromech. Syst. 26, 691–703 (2017)
21. Mohamed Ali, M.S., Abdullah, S.S., Osman, D.C.: Controllers optimization for a fluid
mixing system using metamodeling approach. Int. J. Simul. Model. 8, 48–59 (2009)

You might also like