You are on page 1of 39

This article was downloaded by: [University of Western Ontario]

On: 03 February 2015, At: 10:11


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Heat Transfer Engineering


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/uhte20

Comparing the mixing performance of common types of


chaotic micromixers: a numerical study
ab b b c
Athanasios G. Kanaris , Ioannis A. Stogiannis , Aikaterini A. Mouza & Satish G. Kandlikar
a
Xaar Plc, Cambridge UK
b
Department of Chemical Engineering, Aristotle University of Thessaloniki, Greece
c
Department of Mechanical Engineering, Rochester Institute of Technology, Rochester, NY,
USA
Accepted author version posted online: 19 Nov 2014.

Click for updates

To cite this article: Athanasios G. Kanaris, Ioannis A. Stogiannis, Aikaterini A. Mouza & Satish G. Kandlikar (2014): Comparing
the mixing performance of common types of chaotic micromixers: a numerical study, Heat Transfer Engineering, DOI:
10.1080/01457632.2015.987623

To link to this article: http://dx.doi.org/10.1080/01457632.2015.987623

Disclaimer: This is a version of an unedited manuscript that has been accepted for publication. As a service
to authors and researchers we are providing this version of the accepted manuscript (AM). Copyediting,
typesetting, and review of the resulting proof will be undertaken on this manuscript before final publication of
the Version of Record (VoR). During production and pre-press, errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal relate to this version also.

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
ACCEPTED MANUSCRIPT

Comparing the mixing performance of common types of chaotic micromixers: a

numerical study

Athanasios G. Kanaris1,2, Ioannis A. Stogiannis2, Aikaterini A. Mouza2

and Satish G. Kandlikar3


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

1
Xaar Plc, Cambridge UK
2
Department of Chemical Engineering, Aristotle University of Thessaloniki, Greece
3
Department of Mechanical Engineering, Rochester Institute of Technology, Rochester, NY, USA

Address correspondence to Dr. Aikaterini A. Mouza, Department of Chemical Engineering,

Laboratory of Chemical Process and Plant Design, Aristotle University of Thessaloniki, Greece.

E-mail: mouza@auth.gr

Abstract

High efficiency micromixers are of critical importance in several industrial applications.

Implementing chaotic advection to passively enhance fluid mixing is a well-established

technique. The purpose of the present study is to compare different configurations of passive

mixers. Six different microchannel designs, commonly used in the majority of mixing

applications, were numerically evaluated in the present study. Both two-dimensional and three-

dimensional serpentine-type microchannel designs were investigated. For each design, the effect

1
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

of geometrical parameters was studied using a commercial computational fluid dynamics code.

Results are presented in terms of mixing efficiency and pressure drop. It is revealed that

increasing the flow path by providing compact fluid passages is not the only way to achieve

better mixing; inducing a helical flow along the flow length has an additional benefit in

improving the mixing performance. The results of this study can be used as a basis for further

improving the design of micromixers, while they can also provide the engineers with guiding
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

principles, not for selecting the optimum set of geometrical parameters for a specific

microchannel design but for screening out the least efficient micromixer configurations.

2
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Introduction

Mixing in microfluidic systems is typically dominated by diffusion, rather than turbulence,

due to the laminar flow that usually prevails in the low Reynolds number flows at microscale [1].

Microchannels of different configurations are increasingly being integrated in microfluidic

mixing devices. It is however evident that, when mixing is driven solely by diffusion, long
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

residence time is required. To improve a micromixer performance, either active mixing methods

(e.g. acoustic actuation) or passive mixing features, e.g. surface roughness, complex two-

dimensional or even three-dimensional paths, can be used. Active micromixers - which can

achieve excellent mixing capabilities - are unfortunately difficult to be integrated in micro-

devices and have in general a higher implementation cost, thereby making them unsuitable for

disposable microfluidic devices [2, 3]. For the most appropriate application of passive mixing

features, it is important to understand the flow through these complex channels and assess the

effect of the different geometrical parameters on the mixing performance. This understanding

will then be helpful in the design of robust microfluidic devices.

The design concept of micromixers utilizing chaotic advection is based on the simple idea of

the modification of the channel shape for splitting, stretching, folding and breaking of the flow;

in other words, laminating the flow. A lot of work has already been done in this field. Hong et al.

[2] described the design of a passive mixing microreactor using modified Tesla structure, which

is easy to manufacture and of low cost. A comprehensive review published by Hessel et al. [4]

presents a number of different approaches on passive and active mixing. The authors also

suggest that, as the catalogue of different available micromixer designs is quite broad, it is

3
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

advisable for the future designer to focus mainly on the improvement of the existing designs and

the implementation of additional functions in them, such as heating and sensing. Rawool et al.

[5] suggest that the addition of obstructions as part of the channel geometry can be also

beneficial to micromixer efficiency. Lin and Yang [6] analyzed a planar serpentine microchannel

and presented the flow field and the concentration distribution in the two mixing fluids. Their

study, which was performed for Re=160, suggests that the interfacial area between the fluids is
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

continuously distorted due to the changes in the direction, and the flow trajectories are highly

unstable. Park and Kwon [3] studied a similar serpentine design, extending it to 3D. They

identified rotation, lamination and chaotic advection as the main features that improve mixing in

this type of serpentine micromixers. Tsui et al. [7] placed on the walls of the channels either

grooves or block obstacles that disrupt the flow and suggested that the mixing performance can

be enhanced by having the grooves arranged in a staggered manner, by which the transversal

velocity is largely increased. Jeon and Sin [8] studied, both experimentally and numerically, the

mixing performance of passive planar micro-mixers with different geometries and concluded that

valuable design guidelines can be derived by analyzing the volume fraction of mixing fluids.

Wang and Hu [9] investigated the effect of obstacles in T-type microchannels in the mixing of

fluids. The authors concluded that several geometrical features of these channels as well as

various inlet Re ratios between the inlet branches are not yet fully studied and suggested that

such findings can be useful for providing optimal designs of micromixers. In a more recent

review on micromixers, Kumar et al. [10] discussed different designs and noted that when 3D

curved microchannels are used, mixing can be greatly improved even for very low Re (from

0.001 to 5). Kanaris and Mouza [11] investigated the effect of geometrical parameters by altering

4
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

the original design of a prototype μ-reactor and studied them by comparing Resident Time

Distributions (RTDs) using a numerical approach. The results of their study revealed that the

increase in curvature promotes the mixing performance of the device. In a similar study, Kim et

al. [12] have also numerically investigated the effect of geometrical parameters on the

performance of a T-type micromixer with obstruction elements by calculating the concentration

distribution of two liquids along the channel. Finally, Zhang et al. [13] who have performed
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

simulations on 3D-twisted compression-expansion microchannels by varying the clamping angle

have reported that the mixing performance greatly depends on the angle of each cross-section

turns.

As already mentioned, there are a number of micromixers available in the literature and a

number of studies have focused on optimizing their performance. Mixing efficiency can be

significantly improved by chaotic advection, which can be either generated by special geometries

in the mixing channel (i.e., passive) or induced by an external force (active). While many types

of different designs of passive micromixers have been proposed and investigated by various

researchers, a more detailed study on the effect of parameters describing similar geometrical

features across micromixers channels of different designs is warranted. This paper aims at

comparing the performance of six highly successful micromixing techniques and presenting a

comparison under the same operating conditions. To do this, a common ‗design rule‘ will be

adopted for all of them, in the form of ‗building blocks‘ of similar features. In an effort to

contribute to the designing of more efficient microdevices, the scope of this study is to review

and compare various designs published in the literature. Thus, the effect of the key geometrical

parameters of common ―building blocks‖ of micromixers on the mixing efficiency and on the

5
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

corresponding pressure drop will be investigated. Such evaluation is expected to be of value to

practical designers of micromixers in selecting a specific mixing technique.

Micromixer designs

A great variation of microdevice designs with different geometrical characteristics of the

flow channels have been extensively used as passive micromixers. However, the typical

―building blocks‖ which form the microchannels are rather standard [4]. The most common
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

designs for simple passive micromixers comprise orthogonal blocks, located either on one or two

planes, while curved microchannels, which enhance mixing efficiency by promoting helical

flow, can be also implemented into the device [14]. This study attempts to combine these

fundamental features into representative generic designs. Thus, six designs are chosen and are

illustrated in Figure 1, whereas the edge, W, of the square duct is 100μm, for all cases. Namely

the designs investigated are: a simple T-type straight duct (T) used as reference, a simple

orthogonal serpentine geometry (SO), two orthogonal serpentine designs extending on two

parallel planes with channels resembling the form of the letter C (C) and the letter L (L), and a

curved serpentine extended either on a single plane (SCSP) or on two parallel planes (SCDP).

The length of the fluid path is determined by the repeated mixing elements increased by a T-

type fluid entrance region of constant geometrical characteristics. Four recurrent ―building

blocks‖ are considered for all cases as they were proved to be sufficient for achieving a

quantifiable degree of mixing, while on the other hand are feasible to simulate with the available

computer facilities. Each mixing element is described by two characteristics dimensions (L1, L2)

as illustrated in Figure 2 as well as the channel edge (W) which is set to be constant throughout

this study.

6
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

An initial study was performed for the Y and C mixers. An interesting conclusion drawn is

that the overall mixing efficiency of the device, at least for the range of operating conditions

tested, was not practically affected by the angle of the Y-entrance region; therefore, this

parameter was screened and not included in the design of simulations. Typical mixing elements

are presented in Figure 2. Based on the above, two dimensionless parameters are defined and

used for the parametric design: a length ratio, LR, and a normalized length, ZR. LR is defined as:
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

LR = L1 / L2 (1)

and ZR defined as:

ZR = W / L1 (2)

Since the channel edge, W, is considered constant for this study (100μm), it is obvious that the

length of the device, Lz , is controlled by ZR, while the total flow path, Lt , is controlled by LR. A

characteristic quantity would be the ratio of Lt over Lz, which is an indicator of the level of

compactness of the device. Compactness of such devices has been reported as one of the most

important advantages of microdevices when are used in mixing [10] or heat transfer applications

[15]. The design parameters used, i.e. LR and ZR, are illustrated in Figure 2, while their range is

given in Table 1. It must be noted that, as the microchannel is assumed to have a square cross

section, when the microdevice is considered to be comprised of two planes, the total height of the

microdevice is equal to 2W, as the two planes are assumed to be perfectly attached to each other.

However, as the actual microdevice includes an overhead due to the connection between the two

planes, this might not hold completely true in reality. For the sake of simplicity and for the

purpose of this study, this overhead is considered negligible.

7
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Water and methyl alcohol (CH3OH) were chosen to be the mixing fluids as under normal

conditions they form ideal mixtures and do not react with each other. The thermophysical

properties for both fluids were evaluated at room temperature and considered constant. A similar

approach is used by Kim et al. [12].

In a recent study [11] it is reported that an increase of Re number leads to a better mixing

efficiency. Based on this, in the present study, whose aim is to compare different configurations,
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

the effect of Re has not been taken into consideration. Consequently, for all simulations the flow

rate has the same value that corresponds to laminar flow (i.e., Re has a value of about 2 for both

fluids) to ensure that the prevailing mixing mechanism is the mass diffusion between the liquids.

In the present calculations, the computational fluid dynamics (CFD) code uses the laminar

flow model of ANSYS CFX and the High-resolution Advection Scheme for the discretisation of

the momentum equations. The two fluids, water and methyl alcohol, are considered having

different physical properties and the prevailing mixing mechanism is the mass diffusion between

them. No chemical reaction was included in the model and the fluids are considered

incompressible. Mass-flow boundary conditions are set on the two inlet ports. The flow rate at

each inlet of the Y-entrance was set equal to 10μL/min, which corresponds to Re=1.67 for water

and to Re=1.77 for methyl alcohol, i.e. to laminar flow. An additional flow rate, 50μL/min, was

also used for a group of simulations with the intention to evaluate whether results are

independent of the initial boundary conditions. A pressure boundary condition of atmospheric

pressure is set on the outlet port, while the convergence criterion is the mass-balance residual

value to be less than 10-9. Based on the Schmidt number and fluid properties, the diffusion

coefficient between the two fluids is calculated to be 1.570x10-9m2/s [12].

8
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

In order to compare the mixing efficiency of the different geometries studied it is needed to

quantify the quality of mixing. In the majority of studies, either experimental [16-18] or

numerical [19-21], an Index of Mixing Efficiency (IME) is introduced, based on the standard

deviation of mass fraction from the mean concentration over the cross section area:

  c c 
2
dA

IME  1  A
(3)
 c
2
dA
A
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

In this approach the cross section is divided into several units (in our case the grid elements

of each cross section) and concentration is calculated in each unit. The mass flow averaged value

of the mass fraction of water is calculated using the built-in functions of the CFD code

postprocessing software (ANSYS CFD-Post). Mixing is considered to occur when the water

volume fraction equals methyl alcohol volume fraction (i.e., both attain the value of 0.5). As the

densities of water and methyl alcohol are 997 kg/m3 and 791 kg/m3, respectively, for the case of

water volume fraction of 0.5, the water mass fraction is:

1
c  0.55 .
791
1
997

the mass flows of the mixing fluids are constant over time, the mean concentration c will also

have a constant value of 0.55.

As the numerical diffusion in the CFD calculations can influence the accuracy of the

calculations, a thorough grid dependence study is performed to ensure that the solution is

independent of the grid density. As it has been already stated, trying to keep the computational

demand of the simulation low, only four mixing elements plus the entrance region were tested

for each configuration. As it is expected, only a percentage of the two fluids are actually mixed

9
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

at the outlet of the fourth element. During the grid dependence study, the results were evaluated

in terms of pressure drop, ΔP, and mixing efficiency, IME. All grids were constructed with

structured hexahedral elements and the final maximum size of the element was chosen to be

equal to 3.5μm.

In this study, two different lengths were considered; the total length of the fluid path, Lt, and

the length of the device, Lz (Figure 2). It is apparent that, from the designer's point of view, a
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

certain micromixer is superior to another when it achieves better mixing per unit of actual device

length. Increasing the flow path on the directions perpendicular to the flow direction can be

beneficial to the mixing efficiency, provided that the inevitable increase in pressure drop is also

taken into account, leading to a clear tradeoff.

In order to estimate the performance of this type of micromixers, two response variables are

considered: ΔP/Lz and IME/Lz. The former is defined as the pressure drop per unit of device

length, as proposed above, while the latter refers to the mixing performance, again per unit of

device length. Both of these response variables, for all the designs considered in this study, are

fitted in a second-order approximated model, n, of the generic form:

n  C0  C1ZR  C2 LR  C11ZR2  C22 LR 2  C12 ZRLR (4)

using standard Response Surface Methodology (RSM) [22]. To reduce the number of simulations

needed for each micromixer configuration for obtaining a full factorial design, a Design of

Experiments (DOE) method is applied. This approach has been successfully used before [11, 12].

In order to acquire confidence for the method used and to validate the resulting model, two

additional refining design points were solved for each case.

ACCEPTED MANUSCRIPT
10
ACCEPTED MANUSCRIPT

To assess the overall performance of the designs two performance indices that compare each

microchannel type with the reference case design of a T-type square duct are introduced:


Lt
n  (5)
 
Lt ,Y

IME
Lt
nIME  (6)
IME
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

Lt ,Y

The nΔΡ index is a friction-related term while nIME is a mixing-related term and both can be

used to assess the interplay between mixing enhancement and pressure drop.

Results

CFD code validation

In order to ensure that the selected CFD code can accurately predict fluid flow and mixing

performance in micromixers, the code is validated using experimental data available in the

literature. Liu et al. [23] studied a 3D serpentine microchannel, which consisted of the same

building blocks as the C-type micromixer design model used in this study. In their approach

mixing efficiency is evaluated by observing color variation of a dye (phenolphthalein) using a

CCD camera. Light intensity is used to quantify the uniformity of mixing in a plane visible to the

camera through a transparent ―viewing window‖ machined on the microchannel geometry. A

mixing index, identical to the one used in the present study, was calculated for assessing mixing

performance in three different geometries for various Re numbers.

A C-type model of this work‘s approach is modified in order to meet the geometrical

characteristics of the experimental device [23] and has been then modeled in a series of CFD

11
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

runs using the same model parameters employed throughout this study. Results from the

numerical simulations, shown in Figure 3, are in excellent agreement (deviation less than 3%)

with the experimental results for a range of Re starting from 5 up to 12. Simulations are focused

on the lower part of the experimental Re range due to memory limitations of the available

computational facilities. However, it should be noted that CFD code is validated in a range of Re

numbers significantly higher than the one used thereafter in the study and thus it can be
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

considered that it is valid to predict fluid flow and mixing performance for Re numbers lower

than the range used, where the demand on grid size is also lower.

Results on micromixers comparison

CFD simulations were performed for the six different geometries described in detail in a

previous section and for the range of design variables reported in Table 1. The IME was

evaluated for all studied geometries at the flow outlet using ANSYS CFX Expression Language

(CEL) commands in the ANSYS CFD-Post package.

Using RSM methodology a second-order polynomial equation was fitted to the results, whose

coefficients are presented in Table 2, while the value of R-square is calculated to be over 99% for

all cases. Simulation results show the geometric variations have a distinct influence on mixing.

Figure 4 presents the streamline plots for all designs in a representative area of the conduit,

while Figure 5 shows the mixing efficiency and pressure drop against the total flow path length,

Lt. As it is expected, when the actual device length, Lt, is increased or equally ZR is increased,

both mixing performance and pressure drop are increased.

The characteristic length for all cases is the same (equal to W) and as a result, a linear

increase of pressure drop with unit length can be expected. This trend is indeed shown on Figure

12
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

5a. Based on the above, pressure drop vs. flow path length (ΔP/Lt) is the same regardless of the

microchannel design. This means that, for the low Re values usually encountered in micro-

devices, the number and type of bends does not significantly affect the pressure drop per unit

length. The same trend has been reported by Kumar et al. [10], where for the same Re number in

the laminar regime, friction factor is practically independent of the geometry of the

microchannel.
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

As mentioned above, Lt / Lz is an indicator of the compactness of the device. A valid

estimation of Lt would be:

Lt   L1  L2   number of elements .

ile an estimation of Lz would be:

Lz  L2  number of elements .

ich allows the calculation of their ratio:

Lt
 LR  1 .
Lz

including pressure drop:

P
Lz
 LR  1 .
P
Lt

ads to the prediction of pressure drop vs. the length of the device to be:

P   P    LR  1 (7)
Lz  Lt 

As LR is increased for a given value of ZR, and as pressure drop vs. flow path length (ΔP/Lt)

is found to be constant for this study, an increase of pressure drop per unit of device length

13
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

(ΔP/Lz) is also inevitable. To evaluate the effect of LR, additional simulations for the SO design

were run for a wider range of LR values. The dependence of the micromixer performance on the

value of LR , which defines the channel bend ratio and controls the total fluid flow length, is

presented for the SO type device in Figure 6. It is obvious that the rate of increase of IME

(representing mixing performance) is higher than that of the pressure drop. This might lead to the

conclusion that as LR is increased, the overall performance of the microdevice is improved.


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

However, for the final decision one must rely on the relative importance of mixing efficiency

versus the corresponding increase in pumping cost. For the two-layer configurations (i.e., C-

Type, L-Type and SCDL) the interfacial area between the fluids is distorted and enlarged in a

spiral behavior, effectively promoting the fluid mixing. The effect of channel curvature seems to

be less important than the effect of flow direction change.

The L-Type design performs slightly better in terms of mixing compared to the other designs,

as shown in Figure 5b. In terms of mixing performance, it can be seen that L-type stands out as

the best, albeit with a small difference compared to the C-Type and SCDP, which perform very

similar to each other; SO and SCSP perform lower and also similar to each other (Figure 5b).

The similarity between C-Type and SCDP as well as SO and SCSP can be explained by the fact

that the incorporated designs are practically the same, where SCDP & SCSP include curved

channels and C-Type & SO use orthogonal channels. It must be noted that simulations were also

run for a higher flow rate, i.e. 50μL/min, which corresponds to Re=8.35 and 8.85 for water and

methyl alcohol, respectively, and the results are found to be consistent with the above

conclusion.

14
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

It is also shown that the mixing efficiency per device length (Lz) attains a higher value for the

L-Type mixer. In Table 3 the values of pressure drop and the mixing efficiency of the different

configurations are evaluated for a pair of ZR and LR values and they are compared to the base

case (i.e. T-type design). In all cases, increase in pressure is about 40 to 50%, while the increase

in mixing efficiency is always more than 200%. As mentioned above, for this range of Re

numbers, pressure drop is not notably influenced by the number of bends in the microchannel,
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

but by only its actual length, Lt. This statement is verified also by the practically similar pressure

drop calculated for a typical set of the design variables for all design cases (Table 3). Therefore,

choosing a design for the improvement of the overall performance of the microdevice should

mainly consider the parameter values that improve mixing efficiency.

The question that rises is why L-Type performs better than C-Type in terms of mixing

efficiency per flow path length, for the same LR; it is not the number of bends (since both have

the same) but the induced rotation of the flow that enhances mixing efficiency. The fluid path in

L-Type actually follows a helix, completing a 360 deg turn due to the type of bends introduced

(Figure 7a). On the other hand, in C-type the fluid follows a 3D zig-zag movement (Figure 7b),

which is less effective. As a result, the performance of the micromixer design cannot be

expressed in terms of number of bends, since it also depends on the actual rotation of the fluid

path. The comparison between C-Type and L-Type, and the effect of inducing rotation in the

fluid flow path are also mentioned in the paper by Park and Kwon [3], and the results of this

study agree with their observations.

15
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Conclusions

This numerical study considered five channel geometries as alternatives to the well-known T-

type straight-duct micro-mixer. A CFD software was used to characterize these different

geometries for their effectiveness in mixing and flow impendence. All the tested geometric

designs lead to an increased mixing efficiency in comparison to the typical straight-duct

micromixer. This improvement is characteristically attributed to the presence of channel bends,


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

which create fluid recirculation regions and secondary flow.

 The L-type micromixer tends to have a marginally better performance in terms of mixing

efficiency, since the chaotic behavior of fluid trajectories produces the greatest mixing

intensification. The mixing efficiency for the best case scenario is improved almost 3 times

in comparison to the baseline design. The L-type microdevice design has an additional

advantage over its similar counterpart, the C-type design, as for the same geometrical

design parameters the induced helical flow further improves the performance of the

micromixer.

 The pressure drop of the device is only attributed to the increase in total fluid flow length.

Since the flow is laminar, pressure drop increases linearly with flow length, regardless of

the existence of other geometrical features. This observation allows the designer to seek

only the improvement of mixing efficiency.

 The results from this study provide guidelines for the engineers in selecting appropriate

type of micromixer for a specific application. The micromixers that can be clearly

excluded are identified and guidelines are provided for selecting among the available

choices.

16
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

It can be deduced that the geometries, which induce helical flow path, offer a number of

improvements over typical designs. Further study of these configurations can provide devices

that are more robust. The results of this study can be used as a basis for further improving the

design of micromixers, while they can also provide the engineers with guidelines, not for

selecting the optimum set of geometrical parameters for a specific microchannel design but for

screening out the least efficient micromixer configurations.


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

Nomenclature

A cross-section area, m2

C Parameters of the quadratic model

c mass fraction in each cross section unit (grid element)

c mean concentration over the cross section area, A

IME Index of mixing efficiency

L1, L2 Lengths of specific geometrical features of the microchannels, m

LR, ZR Dimensionless geometrical parameters (Table 1)

L L-type geometry

Lt Total flow length, m

Lz Flow length on the direction of the flow, m

n Approximated quadratic model, used for RSM

nΔP Friction performance index

nIME Mixing performance index

17
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Re Reynolds number

SCDP Serpentine-Curved on a dual plane configuration

SCSP Serpentine-Curved on a single plane configuration

SO Serpentine-Orthogonal geometry

W channel width, m

z direction of flow along channel length, m


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

Greek symbol

ΔP Pressure drop, Pa

18
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

References

[1] Gravesen, P., Branebjerg, J., and Jensen, O. S., Microfluidics-a review, Journal of

Micromechanics and Microengineering, vol. 3, no. 4, pp. 168-182, 1993.

[2] Hong, C.-C., Choi, J.-W., and Ahn, C. H., A novel in-plane passive microfluidic mixer

with modified Tesla structures, Lab on a Chip, vol. 4, no. 2, pp. 109-113, 2004.

[3] Park, J. M., and Kwon, T. H., Numerical characterization of three-dimensional serpentine
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

micromixers, AIChE Journal, vol. 54, no. 8, pp. 1999-2008, 2008.

[4] Hessel, V., Löwe, H., and Schönfeld, F., Micromixers—a review on passive and active

mixing principles, Chemical Engineering Science, vol. 60, no. 8–9, pp. 2479-2501, 2005.

[5] Rawool, A. S., Mitra, S., and Kandlikar, S. G., Numerical simulation of flow through

microchannels with designed roughness, Microfluidics and Nanofluidics, vol. 2, no. 3, pp.

215-221, 2006.

[6] Lin, K.-W., and Yang, J.-T., Chaotic mixing of fluids in a planar serpentine channel,

International Journal of Heat and Mass Transfer, vol. 50, no. 7–8, pp. 1269-1277, 2007.

[7] Tsui, Y.-Y., Yang, C.-S., and Hsieh, C.-M., Evaluation of the mixing performance of the

micromixers with grooved or obstructed channels, Journal of fluids engineering, vol. 130,

no. 7, pp. 0711021-07110210, 2008.

[8] Jeon, W., and Shin, C. B., Design and simulation of passive mixing in microfluidic

systems with geometric variations, Chemical Engineering Journal, vol. 152, no. 2–3, pp.

575-582, 2009.

[9] Wang, C., and Hu, Y., Mixing of liquids using obstacles in Y-type microchannels,

19
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Tamkang Journal of Science and Engineering, vol. 13, no. 4, pp. 385-394, 2010.

[10] Kumar, V., Paraschivoiu, M., and Nigam, K. D. P., Single-phase fluid flow and mixing in

microchannels, Chemical Engineering Science, vol. 66, no. 7, pp. 1329-1373, 2011.

[11] Kanaris, A. G., and Mouza, A. A., Numerical investigation of the effect of geometrical

parameters on the performance of a micro-reactor, Chemical Engineering Science, vol. 66,

no. 21, pp. 5366-5373, 2011.


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

[12] Kim, B. S., Kwak, B. S., Shin, S., Lee, S., Kim, K. M., Jung, H.-I., and Cho, H. H.,

Optimization of microscale vortex generators in a microchannel using advanced response

surface method, International Journal of Heat and Mass Transfer, vol. 54, no. 1–3, pp.

118-125, 2011.

[13] Zhang, Y., Hu, Y., and Wu, H., Design and simulation of passive micromixers based on

capillary, Microfluidics and Nanofluidics, vol. 13, no. 5, pp. 809-818, 2012.

[14] Schönfeld, F., and Hardt, S., Simulation of helical flows in microchannels, AIChE

Journal, vol. 50, no. 4, pp. 771-778, 2004.

[15] Kandlikar, S. G., High Flux Heat Removal with Microchannels—A Roadmap of

Challenges and Opportunities, Heat Transfer Engineering, vol. 26, no. 8, pp. 5-14, 2005.

[16] Jayaraj, S., Kang, S. M., and Suh, Y. K., A review on the analysis and experiment of fluid

flow and mixing in micro-channels, Journal of Mechanical Science and Technology, vol.

21, no. 3, pp. 536-548, 2007.

[17] Bhagat, A. A. S., Peterson, E. T. K., and Papautsky, I., A passive planar micromixer with

obstructions for mixing at low Reynolds numbers, Journal of Micromechanics and

ACCEPTED MANUSCRIPT
20
ACCEPTED MANUSCRIPT

Microengineering, vol. 17, no. 5, pp. 1017-1024, 2007.

[18] Nimafar, M., Viktorov, V., and Martinelli, M., Experimental comparative mixing

performance of passive micromixers with H-shaped sub-channels, Chemical Engineering

Science, vol. 76, pp. 37-44, 2012.

[19] Wu, C. Y., and Tsai, R. T., Fluid mixing via multidirectional vortices in converging-

diverging meandering microchannels with semi-elliptical side walls, Chemical


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

Engineering Journal, vol. 217, pp. 320-328, 2013.

[20] Jeon, W., and Shin, C. B., Design and simulation of passive mixing in microfluidic

systems with geometric variations, Chemical Engineering Journal, vol. 152, no. 2-3, pp.

575-582, 2009.

[21] Alam, A., and Kim, K.-Y., Mixing performance of a planar micromixer with circular

chambers and crossing constriction channels, Sensors and Actuators B: Chemical, vol.

176, pp. 639-652, 2013.

[22] Myers, R. H., Montgomery, D. C., and Anderson-Cook, C. M., Response Surface

Methodology: Process and Product Optimization Using Designed Experiments, Wiley,

2011.

[23] Liu, R. H., Stremler, M. A., Sharp, K. V., Olsen, M. G., Santiago, J. G., Adrian, R. J.,

Aref, H., and Beebe, D. J., Passive mixing in a three-dimensional serpentine

microchannel, Journal of Microelectromechanical Systems, vol. 9, no. 2, pp. 190-197,

2000.

ACCEPTED MANUSCRIPT
21
ACCEPTED MANUSCRIPT

Table 1. Dimensionless parameters used for the study.

Dimensionless parameter Range

ZR=W/L1 0.25-0.50
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

LR=L1/L2 0.30-0.70

22
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Table 2. Response model coefficients for a) ΔP/Lz ; b) IME/Lz.

(a) C L SO SCSP SCDP

C0 70890 77054 69125 107223 143185

C1 -190 -8387 9548 -130058 -169358


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

C2 66841 31513 67166 17362 -26095

C11 -4191 -13691 -15276 128919 149018

C22 4795 43859 5623 5776 23129

C12 -69375 -14339 -114358 -32631 1342

(b) C L SO SCSP SCDP

C0 43.08 165.68 10.16 226.25 300.70

C1 29.57 -530.12 70.15 -326.47 2057.84

C2 -96.27 -374.31 63.40 -123.34 -2343.34

C11 -45.51 578.08 -58.56 276.80 -3026.97

C22 129.13 311.89 -4.72 18.31 2435.92

C12 458.16 1125.28 233.37 116.43 255.75

23
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Table 3. Performance indices of each design for ZR=0.5 and LR=0.3.

Baseline
C L SO SCSP SCDP
(Straight)

ΔP/Lt (Pa/m) 40793 58596 56447 59457 59827 60363


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

nΔΡ 44% 38% 46% 47% 48%

IME/Lt (m-1) 12.1 81.5 100.9 78.5 68.7 53.7

nΙΜΕ 228% 282% 219% 192% 150%

24
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

LIST OF FIGURES

Figure 1. Micromixer designs analyzed in this study.

Figure 2. Design parameters of the microchannels. Also LR = L1 / L2 and ZR = W / L1.

Figure 3. Comparison of CFD code results with experimental ones from Liu et al. [23] using

the Index of Mixing Efficiency.


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

Figure 4. Streamlines inside various micromixer designs: a) SO, b) C, c) L, d) SCSP and e)

SCDP for ZR=0.5 and LR=0.3.

Figure 5. a) Pressure drop and b) Mixing efficiency vs total flow length. Error bars at -5%.

Figure 6. Dependence of the two performance indices on LR values (SO type, ZR=0.24).

Figure 7. Fluid path changes for a) L and b) C mixers.

25
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

T-type straight duct (T) Serpentine-Orthogonal (SO)

C-type (C) L-type (L)

ACCEPTED MANUSCRIPT
26
ACCEPTED MANUSCRIPT

Serpentine-Curved on a single plane configuration Serpentine-Curved on a dual plane configuration

(SCSP) (SCDP)

Figure 1. Micromixer designs analyzed in this study.


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

ACCEPTED MANUSCRIPT
27
ACCEPTED MANUSCRIPT

Lz L2 Lt

L2

L1
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

L1

(a) (b)

Figure 2. Design parameters of the microchannels. Also LR = L1 / L2 and ZR = W / L1

ACCEPTED MANUSCRIPT
28
ACCEPTED MANUSCRIPT
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

error bars at
±3%

Figure 3. Comparison of CFD code results with experimental ones from Liu et al. [23]

using the Index of Mixing Efficiency.

29
ACCEPTED MANUSCRIPT
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

(b)
(a)
ACCEPTED MANUSCRIPT

(c)

30
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

(d) (e)
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

Figure 4. Streamlines inside various micromixer designs: a) SO, b) C, c) L, d) SCSP and e)

SCDP

for ZR=0.5 and LR=0.3.

31
ACCEPTED MANUSCRIPT
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

(b)
(a)
ACCEPTED MANUSCRIPT

32
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

Figure 5. a) Pressure drop and b) Index of Mixing efficiency vs. total flow length.

33
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

Figure 6. Dependence of the two performance indices on LR values (SO type, ZR=0.24).

34
ACCEPTED MANUSCRIPT
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

(a)
ACCEPTED MANUSCRIPT

35
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
Downloaded by [University of Western Ontario] at 10:11 03 February 2015

(b)

Figure 7. Partial view of the fluid path focusing on the directional changes for a) L and b) C

mixers.

36
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Athanasios G. Kanaris holds a diploma in Chemical Engineering and a

PhD on the numerical and experimental study of plate heat exchangers for

their optimal design. He has also worked as a post-doctoral researcher in

the Department of Industrial Energy of the Ecole des Mines de Douai,


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

France, optimizing the design of fin and tube heat exchangers. He is an Associate Member of the

Institute for Chemical Engineers and he currently serves as a Fluid Systems Design Engineer in

Xaar plc working on the design of complex fluidic paths in the microscale and hydraulic

systems.

Ioannis A. Stogiannis is a Ph.D. candidate in the Chemical Engineering

Department of the Aristotle University of Thessaloniki. He holds a diploma

degree in Chemical Engineering from the Aristotle University of

Thessaloniki and a M.Sc. in Advanced Chemical Process Design from The

University of Manchester. His research interests include heat transfer, micro equipment design

and CFD.

Aikaterini A. Mouza is an Assistant Professor in the Department of

Chemical Engineering of the Aristotle University of Thessaloniki (AUTh),

Greece. She holds a Diploma and a PhD in Chemical Engineering from

AUTh. Her main research interests include conventional and micro-

37
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

equipment design, process simulation, CFD, use of nanofluids for heat transfer enhancement and

multiphase flow. Using advanced non-intrusive measuring techniques (μ-PIV, LIF, high speed

photography) she studies the hydrodynamic behavior of micro equipment.

Satish G. Kandlikar is the Gleason Professor of Mechanical Engineering at


Downloaded by [University of Western Ontario] at 10:11 03 February 2015

RIT. He received his Ph.D. degree from the Indian Institute of Technology in

Bombay in 1975 and has been a faculty there before coming to RIT in 1980.

He has worked extensively in the area of flow boiling heat transfer and CHF

phenomena at microscale, single-phase flow in microchannels, high heat flux chip cooling, and

water management in PEM fuel cells. He has published over 200 journal and conference papers.

He is a Fellow member of ASME and a former Associate Editor of ASME Journal of Heat

Transfer. He has received the RIT’s Eisenhart Outstanding Teaching Award in 1997 and Trustees

Outstanding Scholarship Award in 2006. He has received the 2008 Rochester Engineer of the

Year award from Rochester Engineering Society. He is the recipient of the 2012 ASME Heat

Transfer Memorial Award. Currently he is working on DOE and GM sponsored projects on Fuel

Cell water management under freezing conditions, and an NSF sponsored projects on developing

nanostructures for enhanced pool and flow boiling.

38
ACCEPTED MANUSCRIPT

You might also like