You are on page 1of 26

Sensors and Actuators B 259 (2018) 677–702

Contents lists available at ScienceDirect

Sensors and Actuators B: Chemical


journal homepage: www.elsevier.com/locate/snb

Recent advances and applications of micromixers


Chia-Yen Lee a,b , Lung-Ming Fu b,c,∗
a
Department of Vehicle Engineering, National Pingtung University of Science and Technology, Pingtung 912, Taiwan
b
Graduate Institute of Materials Engineering, National Pingtung University of Science and Technology, Pingtung 912, Taiwan
c
Department of Biomechatronics Engineering, National Pingtung University of Science and Technology, Pingtung 912, Taiwan

a r t i c l e i n f o a b s t r a c t

Article history: Micromixers are crucial components within micro biomedical systems. This article presents a compre-
Received 10 September 2017 hensive review of the main state-of-the art applications of micromixers in biomedical systems over
Received in revised form the past ten years. The article commences by reviewing the fundamental fluidic behaviors involved in
27 November 2017
microfluidic mixing. The biomedical applications of micromixers are then described in regard to their
Accepted 4 December 2017
Available online 18 December 2017
use particularly for (1) sample concentration, (2) chemical synthesis, (3) chemical reaction, (4) polymer-
ization, (5) extraction and purification, (6) biological analysis, and (7) droplet/emulsion processes and
others. The review concludes with a brief statistical analysis of the published literature in the micromixer
Keywords:
Biomedical application field and an overview of the relative advantages of passive micromixers compared to active micromixers.
Microfluidic mixing © 2017 Elsevier B.V. All rights reserved.
Micromixer

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 677
2. Sample concentration [16–47] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 679
3. Chemical synthesis [48–88] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 683
4. Chemical reactors [89–128] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 686
5. Polymerization process [129–158] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687
6. Extraction and purification processes [159–191] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 689
7. Biological analysis processes [192–231] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 691
8. Droplet/emulsion and other processes [232–260] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 693
9. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 696
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
Biography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702

1. Introduction
result in laminar flow, and hence species mixing occurs mainly
as a result of diffusion, which is an inherently slow process. Con-
Microfluidic devices have had a significant impact on the field
sequently, efficient microfluidic mixing schemes are required to
of biomedical diagnostics, and are widely used throughout the
increase the throughput of biomedical microsystems and make
drug development, drug delivery and bio-medical research indus-
possible the development of micro-total-analysis systems (␮TAS)
tries [1]. The diminutive scale of the flow channels in microfluidic
[4–6] and lab-on-a-chip (LOC) devices [7,8].
systems increases the surface-to-volume ratio, and is thus advan-
The Reynolds number, Re, and Peclet number, Pe, are two
tageous for many biomedical applications [2,3]. However, the very
of the most crucial parameters when designing and evaluating
low Reynolds number regimes produced in such microchannels
micromixers. The Reynolds number represents the ratio of the fluid
momentum to the viscous friction force, and is defined as
∗ Corresponding author at: Graduate Institute of Materials Engineering, Pingtung Re = VLh /, (1)
912, Taiwan.
E-mail addresses: loudyfu@mail.npust.edu.tw, loudyfu@yahoo.com.tw where V is the average flow velocity, Lh is the hydraulic diameter
(L.-M. Fu). and  is the kinematic viscosity. At the macroscale, turbulent flow

https://doi.org/10.1016/j.snb.2017.12.034
0925-4005/© 2017 Elsevier B.V. All rights reserved.
678 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

Fig. 1. Schematic illustration of microfluidic cell culture device for investigating cell viability under different glucose concentration gradients. (Reprinted from Ref. [16] with
permission of Springer.).

Fig. 2. (a) Optical micrograph of microfluidic system, (b) Electrokinetic preconcentration mechanism at micro/nanochannel junction, (c) cross-sectional view of section
A-A’ in (b) showing overlapping EDL in nanochannel, and (d) variation over time of average fluorescence intensity within region of interest (shown by rectangle in inset).
(Reprinted from Ref. [19] with permission of Springer.).

is indicated by Reynolds numbers with a value of around Re ∼2000 proposed a discussion of the operation conditions in more detail for
in pipe flow. However, at the microscale, turbulent flow is seldom micromixer between the Reynolds number Re and Peclet number
achieved, and most applications involve laminar flow with a low Pe.
Reynolds number. Mixing methods commonly depend on the generation of chaotic
The Peclet number represents the ratio of mass transport due to advection, in which the fluid motion varies irregularly and thus
convection to that due to diffusion, and is expressed as causes quantities such as the pressure and velocity to vary ran-
domly in both space and time. Some important review literatures
Pe = VL/D, (2)
[10–12] interpreted the microfluidic mixing mechanisms and
where V is again the average flow velocity, L is the length of the mix- methods of the micromixer in detail. Chaotic advection can be
ing path and D is the diffusion coefficient [9]. Nguyen and Wu [10] generated by stirring the flow, and is highly effective at low
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 679

Fig. 3. Diffusive gradient mixer used to perfuse cells cultured in downstream microchambers with different concentrations of drugs either sequentially or in combination.
(Reprinted from Ref. [21] with permission of Royal Society of Chemistry.).

Fig. 4. Bacteria growth in oxygen gradient-generating microfluidic platform fabricated using multi-layer soft lithography technique. (Reprinted from Ref. [24] with permission
of Royal Society of Chemistry.).

Reynolds numbers due to the resulting splitting, stretching, folding respectively. The mixing index ranges from 0 (0%, fully unmixed)
and breaking up of the species streams. However, besides chaotic to 1 (100%, fully mixed).
advection, molecular diffusion also plays a key role in accomplish- The following sections of this paper present a detailed review
ing mass transfer in micromixers. In fact, in flow regimes where the of seven of the most common applications of micromixers in
flow is strictly laminar, species mixing occurs almost entirely as a biomedical systems, namely (1) sample concentration, (2) chemical
result of molecular diffusion between layers of different concentra- synthesis, (3) chemical reaction, (4) polymerization, (5) extraction
tions. Under such conditions, it is easier to achieve a more effective and purification, (6) biological analysis, and (7) droplet/emulsion
mixing performance if the thickness of the fluid layers is less √ than processes and others.
the characteristic diffusion length (generally approximated as Dt,
where D is the diffusion coefficient and t is the residence time of 2. Sample concentration [16–47]
the species). It is thus desirable to design the microchannel in such
a way as to increase the contact surface area between adjacent fluid Concentration gradients of diffusible substances play an impor-
layers; thereby minimizing the length of the diffusion path between tant role in biological pattern formation and angiogenesis [16,17].
them. However, despite the importance of concentration gradients in
When evaluating the degree of mixing at a cross-section in a biology, relatively few techniques are available for generating and
mixing channel, the mixing index, M, can be quantified as [13–15]. maintaining gradients in low fluid flow rate regimes. Jiang et al.
[18] constructed a microfluidic chip in which diffusive mixing was
⎛  ⎞
 N 
 1  C − C̄
enhanced by a chaotic advection effect generated by an asymmetric
M = ⎝1 −  ⎠ × 100%
i herringbone structure patterned in the ceiling of the microchan-
(3)
N C̄ nel. The experimental results showed that the chip enabled both
i=1
the composition and the shape of the immobilized gradient to be
precisely controlled.
where N is the total number of sampling points, and Ci and C̄ are the Yeh et al. [16] fabricated a polydimethylsiloxane (PDMS) con-
normalized concentration and expected normalized concentration, centration gradient microfluidic chip for determining the optimal
680 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

Fig. 5. Schematic representation of chemical reaction circuit used for parallel screening of in situ click chemistry library. (Reprinted from Ref. [55] with permission of Royal
Society of Chemistry.).

glucose concentration in endothelial cell (EC) viability tests (Fig. 1). Anwar et al. [19] fabricated an integrated micro-nano-fluidic
In the proposed device, the adhesive/non-adhesive areas of the system for the mixing and preconcentration of dissolved proteins.
EC detection zone were patterned using a micro-contact printing The system consisted of an integrated micromixer (IMM) based on
(␮CP) method and the glucose concentration gradient was adjusted an unbalanced split and cross-collision of the fluid streams and a
by changing the flow rate. It was found that the optimal glucose preconcentrator with nanochannels formed by the electrical break-
concentration for EC viability tests ranged from 8.66–15%. down of a PDMS membrane (Fig. 2(a)). Fig. 2(b) illustrates the basic
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 681

Table 1
Summery of micromixer for concentration applications.

Ref. and First Year Micromixer type Flow rate Index (%) Objective Materials
author

[28] Schertzer 2012 Dielectric mixer N/R M: 99.7% Particle concentration AP, pNPP
(Active) detection
[29] Hong 2010 Circular mixer Re = 4 M: 95.8% Methanol Methanol, MOX
(Passive) concentration
detection
[30] Hou 2011 Circular mixer Re = 8–14 M: 92.5% Glucose concentration Glucose, DNS
(Passive) detection
[31] Wang 2012 Circular mixer Re = 0.5–10 M: 95% Methanol Methanol, MOX
(Passive) concentration
detection
[32] Pang 2013 Laminar mixer 20–120 ␮L/min N/R Gaseous carbonyl GLY, PFBHA
(Passive) analysis
[33] Lok 2011 SHM mixer (Passive) 1.67–16.7 ␮L/min M: 90% Luminol-peroxide Luminol, H2 O2 /Co(II)
detection
[34] Fu 2013 Circular mixer Re = 0.5–10 M: 95% CH2 O concentration CH2 O, Fuchsine
(Passive) detection
[35] Chiu 2013 EOF mixer (Active) Applied voltage 50 V M: 90% Preconcetration Rh B, Rh 6G, HCl
[36] La 2013 CSM mixer (Passive) Re = 42 M: 90% Colorimetric analysis Phenolphthalein,pH
indicator
[37] Javanmard 2014 Chaotic mixer (Passive) 18–60 ␮L/h M: 95% Sample preparation Beads, Blood
[38] Fu 2014 Circular mixer Re = 0.5–8 M: 95% CH2 O concentration CH2 O, Fluoral-P
(Passive) detection
[39] Fu 2014 Vortex mixer (Passive) Gas pressure driving M:100% SO2 and CH4 O SO2 , CH4 O, Methylene
force of 1.5 kg/cm2 concentration blue, MOX
detection
[40] Abed 2016 Serpentine mixer 0.2–24.8 mL/min N/R Heat transfer and W/GLY, W/SUC
(Passive) pressure-drop
measurements
[41] Floque 2016 N/R 30–150 ␮L/min M: 95% Boron concentration Boron, sulfuric acid
detection
[42] Ibarlucea 2016 Tesla mixer (Passive) 10 ␮L/min N/R Glucose concentration Glucose, GOx, HRP,
detection ABTS
[43] Gonzalez 2016 Spiral mixer (Passive) Driving force by N/R Glucose concentration Glucose, GOx, HRP, KI
capillarity detection
[44] Anh 2016 Serpentine mixer 0.1–10 mL/h M: >90% Carcinoembryonic CEA, SEB
(Passive) antigen detection
[45] Wang 2016 Chaotic mixer (Passive) 50 ␮L/min M: >90% Isothermal titration BaCl2, RNase A, 2 CMP
calorimetry
[46] Kuo 2017 Square-wave mixer 7.3 mm/s M: 76% Blood plasma mixing Blood plasma, DI water
(Passive)
[47] Kuo 2017 Square-wave mixer 400 rpm (CD mixer) M: 91.4% Blood plasma mixing Blood plasma, DI water
(Passive)

N/R: not report; M: Mixing index.

operation of the ion-selective protein concentration mechanism Notably, the device contained only one mixing stage, and was thus
within the proposed device [20]. Moreover, Fig. 2(c) shows the for- relatively small and easily integrated with LOC platforms. Friedrich
mation of an overlapping electrical double layer (EDL) within the et al. [23] presented a surface plasmon resonance (SPR)-based
nanochannel, which effectively creates an ion-selective membrane biosensor with an integrated microfluidic concentration gradi-
and hence produces a concentration effect. The preconcentra- ent generator for the quantitative assessment of ovarian cancer
tion performance of the proposed device was evaluated using an biomarkers. The biosensor was used to detect the immobilization
FITC-BSA sample; with the measured fluorescence intensity being of ovarian cancer marker antibodies and to quantify the concen-
compared with that of standard solutions with concentrations of tration of r-PAX8. The results showed that the device achieved a
50 ␮M and 100 ␮M, respectively (Fig. 2(d)). detection limit of around 5 nM and a dynamic range with around 1
Kim et al. [21] presented a programmable microfluidic cell array order of magnitude.
for combinatorial drug screening (Fig. 3). The device incorporated Polinkovsky et al. [24] presented two PDMS-based microfluidic
64 individually addressable cell culture chambers, in which the cells devices capable of generating a series of nine gas mixtures with dif-
were exposed either sequentially or simultaneously to 64 pair-wise ferent oxygen concentrations by means of a gas-mixing effect. Both
concentration combinations of two drugs. The cell culture array was devices comprised a two-layer structure, i.e., a flow layer for liquids
used to screen and optimize various combinatorial drug treatments and cell cultures and a gas layer for gas mixtures (Fig. 4). Moreover,
for PC3 prostate cancer cells. However, the authors stated that the in both cases, the two layers were separated by a thin PDMS mem-
proposed device could be equally applied to identify combinatorial brane, which allowed molecular diffusion to take place from the
drug treatments for a much wider variety of diseases. gas layer to the flow layer and produced a subsequent change in
Friedrich et al. [22] developed a novel microfluidic concen- the absorbed oxygen concentration of the liquid/cell culture as a
tration gradient generator capable of producing either linear result. The two devices were designed in such a way as to produce
or exponential concentration ranges. The device consisted of a nitrogen-oxygen mixtures with oxygen concentrations varying lin-
microfluidic channel with two inputs and a single obliquely-angled early between 0 and 100% (Device 1) and exponentially between 0
surface groove micromixer within the base of the channel to induce and 20.9% (Device 2), respectively. In both cases, the devices were
secondary flows and create the required concentration gradient. designed to facilitate cell culturing in semi-permeable microcham-
682 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

Fig. 6. (a) Silicon-based SU8 master mold for PDMS synthesis chip. (b) Mixing effect within first two mixing chambers of micromixer. (Reprinted from Ref. [59] with permission
of Elsevier.).

Fig. 7. Photographs of: (a) X-micro mixer (LTF), (b) micro residence time module (LTF), and (c) PTFE tube reactor. (Reprinted from Ref. [60] with permission of Elsevier.).

Fig. 8. (a) Geometry of multi-inlet vortex mixer, and (b) MoS2 nanoparticle synthesis. (Reprinted from Ref. [61] with permission of Elsevier.).

bers connected to the flow channel. In general, the experimental demonstrated by culturing three bacteria samples with different
results revealed that the division rate of bacteria exposed to a oxygen requirements, namely E. coli, A. viscosus and F. nucleatum.
nitrogen-oxygen gas mixture with a 12% oxygen concentration was Toh et al. [26] developed a microfluidic 3-D gradient-generating
approximately 5 times higher than that of bacteria exposed to a multiplexed device for studying drug hepatotoxicity. The proposed
mixture containing no oxygen. Lam et al. [25] developed a PDMS device contained a multiplexed microchannel consisting of cell
microfluidic device incorporating a differential oxygenator for the culture compartments and medium perfusion compartments to
culturing of anaerobic and aerobic species. In the proposed device, perform the simultaneous screening of multiple dose-dependent
a mixed nitrogen-oxygen gas was flowed through a microchannel drug responses. The hepatotoxicity analysis results obtained for
network and used to regulate the dissolved oxygen concentration five model hepatotoxic drugs (i.e., acetaminophen, diclofenac,
in an adjacent channel filled with culture media via a process of quinidine, rifampin and ketoconazole) showed that drug-based
molecular diffusion. The feasibility of the proposed device was hepatotoxicity is strongly affected by the drug metabolism. It was
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 683

Table 2
Summery of micromixer for chemical synthesis applications.

Ref. and First Year Micromixer type Flow rate Index (%) Objective Materials
author

[69] Daito 2006 K-M mixer (Passive) Re = 200–800 M: 90% Bisphenol F synthesis Bisphenol F, HBA
[70] Luty 2011 SIMM-V2 mixer 0.1–10 mL/min N/R Gold nanoparticles HAuCl4 , NaBH4
(Passive) synthesis
[71] Jung 2012 SAR mixer (Passive) Re ∼50 M:∼ 90% Recrystallization of LDPE, 1-dodecanol
polyethylene
submicron particles
[72] Kumar 2012 4-way mixer (Passive) 35–1000 ␮L/min N/R Ag nanoparticles AgNO3 , KOH, oleic acid
synthesis
[73] Yamamoto 2013 T-mixer (Passive) Re ∼2000 N/R ZIF-8 nanoparticles ZIF-8, Zn2+, 2-MeIM
synthesis
[74] Haroun 2013 Full-loop mixer 2–15 mL/min N/R [11 C]raclopride DMR, [11 C]Mel
(Passive) synthesis
[75] Ono 2014 T-mixer (Passive) 5–75 g/min N/R Ca1−x Srx TiO3 Ca(NO3 )2 , TiO2,
nanoparticles synthesis Sr(NO3 )2 , NaOH
[76] Yang 2015 Tesla mixer (Passive) Re = 0.1–100 M: 94% Immunofluorescence H1975 cancer cells,
analysis L858R
[77] Deng 2015 T-mixer (Passive) 333–667 mL/min M: 99% 2,4,5-trifluorobenzoic EtMgBr,
acid synthesis 2,4,5-Trifluoro-
bromobenzene
[78] Maeki 2015 SHM mixer (Passive) 0.1–1 mL/min M: >95% Lipid nanoparticles POPC, ethanol, NaCl
synthesis
[79] Santana 2015 3-types T-channel Re = 1–160 M: 99% J. curcas oil conversion J. curcas oil, ethanol
mixer (Passive)
[80] Thiele 2015 T, SAR, and Dean flow Re = 27–457 M: 93% Silver seed particles TSC, PSSS, NaBH4 ,
mixer (Passive) synthesis AgNO3
[81] Veldurthi 2015 Circular mixer (Active) Re = 0.006 M: 90% RIF loaded TiO2 RIF, TiO2
synthesis
[82] Kim 2016 T-shape mixer 4.5 mL/min M:80% Y: 91% Organic synthesis PhLi, CICO2 Me, Bu3 SnCI
(passive)
[83] Baek 2016 SHM mixer (Passive) Re = 1.76–10.6 M: 90% PDA nanoparticles PCDA, DMSO, ␣-CD
synthesis
[84] Thiele 2016 Dean flow mixer 13.3–41.7 ␮L/s M: ∼100% Au nanoparticles CTAC, HAuCl4 , NaBH4
(Passive) synthesis
[85] Santana 2017 T-mixer with static Re = 100 M: 99% Biodiesel synthesis Sunflower oil, ethanol,
elements (Passive) NaOH
[86] Thiele 2017 Dean flow mixer Re = 5–350 M: 100% (Re >250) Au nanoparticles CTAC, HAuCl4 , NaBH4
(Passive) synthesis
[87] Pessoa 2017 D, D-bends, D-barriers 25–90 ␮L/min M: 85% Chitosan/ATP CHI, ATP
mixer (Passive) nanoparticles synthesis
[88] Thiermann 2017 SFIMM mixer (Passive) 0.2–2.4 mL/min N/R Block copolymer PB130 -b-PEO66 H,
synthesis acetic acid

N/R: not report; M: Mixing index; Y: Yield efficiency.

also shown that the device was more sensitive to drug-mediated be enhanced simply by adjusting the diffusion distance or strength-
hepatotoxicity than a conventional multi-well plate. ening the convection effects in the flow mixing regions.
Yan et al. [27] developed a two-layer microfluidic slipchip-based Kim et al. [53] developed a PDMS-based micromixer to facili-
reaction microarray in which the sample and reagent formed two tate the reaction of glucose and glucose oxidase (GOX) in which
orthogonal concentration gradients, respectively, and mixing was the mixing channel was patterned with an alligator tooth struc-
achieved by sliding the two layers of the device relative to one ture designed to produce an alternate stretching and folding of the
another. The performance of the proposed device was investigated species stream. The mixing index was found to range from 81–92%
by injecting solutions of 45 nM ˇ-gal and 45 ␮M FDG (Fluores- for Reynolds numbers of Re = 0.08–16. Moreover, a mixing index
cein di-ˇ-d-galactoside) into the upper and lower layers of the of more than 70% was obtained within a mixing distance of less
device, respectively. The reaction products were found to be con- than 2.3 mm. Wang et al. [54] presented a microfluidic chemical
sistent with those obtained in a 384-well plate; thereby confirming reaction device consisting of a nanoliter mixer, a chaotic mixer, a
the reliability of the platform as a tool for investigating the enzy- microfluidic multiplexer and a 32-microvessel array integrated on
matic reactions of ˇ-gal. Table 1 presents a brief comparison of the a single microfluidic platform (Fig. 5). It was shown that the pro-
micromixer-based concentration systems reported in [28–47]. posed device enabled up to 32 click reactions to be performed in a
parallel manner. In a later study [55], the same group developed a
2nd-generation device capable of performing 1024 in situ click reac-
tions in parallel using a bovine carbonic anhydrous II click system.
3. Chemical synthesis [48–88]
Compared to the first device, the 2nd-generation device reduced
the time required to prepare a single click reaction from 1 min to
Chemical synthesis in microchannel networks has attracted sig-
17 s. Furthermore, the use of SPE purification and electrospray-
nificant interest in the literature over the past few decades [48–52].
ionization mass spectrometry with multiple reaction monitoring
Notably, the use of microfluidic platforms for chemical synthesis
(ESI-MRM) enhanced the hit sensitivity of the device and there-
has benefits which extend beyond the laboratory and into industry
fore enabled a significant reduction to be made in the enzyme and
since microfluidics increases the yield and selectivity of the synthe-
reactant consumption.
sis process through more rapid mixing and an improved thermal
management. Furthermore, the efficiency of the mixing process can
684 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

Fig. 9. (a) Experimental setup for recipitation of nanoparticles, and (b) photograph of applied microfluidic T-mixer. (Reprinted from Ref. [95] with permission of Elsevier.).

Fig. 10. Schematic illustration of reaction process in high-pressure, high-temperature microfluidic water system. (Reprinted from Ref. [97] with permission of John Wiley
and Sons.).

Fig. 11. Experimental setup and geometric design of SAR micromixer in microreactor. (Reprinted from Ref. [98] with permission of Elsevier.).

Kawasaki et al. [56] presented a two-step, high-temperature first mixing process was designed to produce a supercritical aque-
micromixing process for the synthesis of anatase TiO2 nanopar- ous KOH solution from supercritical water (SC-H2 O) and KOH, while
ticles with a controlled size and a highly crystalline structure. The the second mixing process combined this KOH solution with aque-
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 685

suppressing secondary nucleation and growth. Chung et al. [59]


developed a simple microreactor based on short-mixing-length
baffles for the synthesis of silica nanoparticles (Fig. 6). The simula-
tion and experimental results showed that the use of three mixing
units with a gap size of 50 ␮m was sufficient to achieve a mix-
ing index of more than 90% via molecular diffusion effects in the
low Reynolds number regime (Re < 0.1) or chaotic and convection
effects in the high Reynolds number regime (Re > 40). In addition,
it was shown that the size of the synthesized particles could be
controlled by tuning the flow rate, reaction time and reaction tem-
perature, respectively.
Baumgard et al. [60] developed two microstructured devices for
the continuous synthesis of platinum (Pt) nanoparticles consisting
of X-type mixers and two different residence time modules, namely
a micro residence time module and a PTFE tube reactor, respectively
(Fig. 7). In the proposed devices, Pt-salt solution was first mixed
with ethylene glycol in the X-type micromixer to produce Pt-seeds
and the resulting acidic solution was then flowed to the residence
Fig. 12. Split-and-recombine micromixer used to create multiple layers of mixtures time module to prompt Pt nanoparticle synthesis. To adjust the
of cyclohexanone oxime and cyclooctane and oleum and ε-caprolacta. (Reprinted temperature during the synthesis process, the residence time mod-
from Ref. [105] with permission of John Wiley and Sons.).
ules were immersed in thermostats (Huber Polystat CC3) filled with
silicone oil. The experimental results showed that the proposed
two-step process allowed for the preparation of nanoparticles with
ous Ti(SO4 )2 to form TiO2 nanoparticles. It was shown that the a tunable particle size and a narrow size distribution. Bensaidet
average particle size could be controlled in the range of 13–30 nm et al. [61,62] developed a multi-inlet vortex microfluidic mixer for
by adjusting the concentration of the KOH solution used in the first the synthesis of molybdenum sulfide nanoparticles (Fig. 8). The per-
mixing process. Li et al. [57] developed a micro-continuous flow formance of the proposed device was evaluated under both laminar
method for the synthesis of ZnO nanoparticles by first mixing the and turbulent flow conditions by means of experiments and CFD
precursor solution using a static mixer and then injecting the mix- simulations. The results showed that nanoparticles with average
ture into a carrier stream. The feasibility of the proposed method diameters of 135 nm, 84 nm and 64 nm were obtained given inlet
was demonstrated using Zn(Ac)2 and NaOH as reactants and binary flows with Reynolds numbers of Re = 832, 4160 and 8320, respec-
solvents of water and ethylene glycol. The results showed that tively. In addition, it was shown that an optimal mixing of the
ZnO nanoparticles with a homogeneous size distribution could be reactants was obtained for a Reynolds number of Re = 8320.
obtained given the use of a binary solvent with a water content of Santana et al. [63] conducted a numerical investigation into
less than 50 v%. The high homogeneity was attributed to a rapid the mixing and reaction of Jatropha curcas oil and ethanol for
mixing of the reactants and solvent in the static micro-mixer and the synthesis of biodiesel in three different micromixers, namely
the fast heat transfer produced within the small droplets of the a T-micromixer, a Cross-micromixer and a Double-T-micromixer.
reaction mixture as a result of a flow-induced segment-internal For each type of device, the simulations investigated the fluid
convection effect. mixing behavior at different Reynolds numbers and the oil con-
Kawasaki et al. [58] presented a continuous supercritical version efficiency for different Reynolds numbers and residence
hydrothermal method based on a T-shaped microfluidic mixer for times, respectively. The results showed that the T-micromixer,
the synthesis of NiO nanoparticles. It was shown that the size of Cross-micromixer and Double-T-micromixer achieved maximum
the synthesized particles reduced from 54.3 to 20.1 nm as the inner conversion efficiencies of 98.48%, 98.55% and 96.00%, respectively.
diameter of the T-shaped mixer was decreased from 2.3 to 0.3 mm Tian et al. [64,65] developed a simple approach for the continu-
and the flow rate was increased from 30 to 60 g/min. The experi- ous synthesis of high quality CdSe quantum dots (QDs) in which
mental and CFD simulation results revealed that the reduction in the nucleation intensity of the QDs was enhanced by preheating
the particle size was the result mainly of a higher heating rate, the Cd and Se precursors in an oil bath prior to their mixing in a
which increased the primary nucleation rate while simultaneously

Fig. 13. Characteristics of five micromixers and four types of flow regime observed in different micromixers. (Reprinted from Ref. [107] with permission of Elsevier.).
686 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

Fig. 14. Microreaction system with multilayer passive micromixer used for polymerization of amino acid NCAs. (Reprinted from Ref. [132] with permission of Royal Society
of Chemistry.).

a step-by-step rapid mixing and heating approach for inducing the


Sonogashira coupling reaction without the need for organic sol-
vents or specific ligands (Fig. 10). It was shown that the desired
products could be obtained in quantitative yields within approx-
imately 4 s when using a ligandless PdCl2 catalyst in pure water
at 16 MPa and 250 ◦ C. Moreover, exceptionally high turnover fre-
quencies of up to 4.3 × 106 h−1 were observed. Notably, the desired
products remained on the surface of the aqueous solution, while
the catalyst was deposited as Pd0 . As a result, the products could
be easily isolated using either a phase separation or filtration tech-
nique.
Chen et al. [98] presented a microreactor with a split-and-
recombine (SAR) structure for the synthesis of imidazole-based
carbohydrate derivatives (Fig. 11). The SAR structure induced a
combined diffusion and chaotic advection effect, and therefore
enhanced both the mixing index and the reaction performance. The
experimental results showed that the proposed device achieved
a similar yield to that of a macroflask, but with a far faster syn-
thesis time. In a later study [99], the same group developed a
Fig. 15. Schematic illustration of fast polycondensation process in continuous-flow SAR ␮-reactor for mixing fluids at Reynolds numbers ranging from
microreactor with micromixer. (Reprinted from Ref. [134] with permission of John
Wiley and Sons.).
Re = 0.01–100 and viscosities of  = 0.000855–0.186 kg m−1 s−1 . The
simulation results showed that the SAR structure induced a 3-
D rotational flow, which prompted an intense stretching of the
micromixer. It was shown that the preheating process enabled the species contact interface and improved the mixing performance
synthesis of CdSe QDs with a smaller size and a narrower size dis- as a result. The 3-D structure of the device enabled the two fluids
tribution. Similar microfluidic devices have also been proposed for to generate sequential stretching and folding, increasing the inter-
the synthesis of ginkgolide B, ZIF-8 and HRPIBs [66–68]. Table 2 facial area between two fluids and, thus, facilitating mixing [100].
presents a brief overview of the proposals presented in [69–88] for Mozharov et al. [101] proposed an improved method for deriving
the use of micromixers in chemical synthesis applications. kinematic information about the reaction process in microfluidic
systems based on a step change in the flow rate and the applica-
tion of real-time noninvasive Raman spectrometry measurements
4. Chemical reactors [89–128]
at the end of the flow line. Taking the Knoevenagel condensation
reaction for illustration purposes, it was shown that the proposed
Mixing is a highly important unit operation in process indus-
system offered several important advantages over conventional
tries, and hence the academic study of mixing in chemical reactors
methods, including a fivefold reduction in the time required to
is of great practical significance. Micro-mixing works simultane-
obtain the kinematic data and a tenfold reduction in the reagent
ously with macro-mixing in the chemical reactors used in process
consumption.
industries, and plays a key role in achieving the optimal selectivity
Han et al. [102] performed CFD simulations based on a standard
with respect to the desired products [89–93]. Micromixers also play
engulfment (E) model and a finite-rate/eddy-dissipation (FR/ED)
a critical role in determining the performance of the microreactors
model to investigate the micromixing performance of viscous fluids
used for the manipulation of reagents or catalyst concentrations in
in a stirred tank reactor fitted with a Rushton turbine. The valid-
the chemical processing field.
ity of the numerical model was validated using experimental data
Schwarzer et al. [94–96] conducted numerical and experi-
reported in the literature. The simulation results showed that the
mental investigations into the precipitation of barium sulfate
micromixing performance and reaction rate both improved with a
nanoparticles in a high capacity microfluidic T-mixer operating
higher agitation speed, a lower fluid viscosity, and a feeding loca-
in the high Reynolds number regime (Re = 200–15000) (Fig. 9).
tion closer to the discharge area of the rotating impeller. Duan et al.
The simulations were based on a population balance equation
[103] used a combined CFD and E-model approach to simulate the
and a global mixing model. The results provided many useful
micromixing effect in single-phase stirred tanks with parallel com-
insights into the micromixing, nucleation and growth stages of
petitive reaction systems. In the proposed model, the turbulence
the precipitation process. Moreover, a good general agreement
kinetic energy and dissipation in the reaction zone were obtained
was observed between the predicted size of the precipitated par-
by solving the transport equations of the mixture fraction and its
ticles and that observed experimentally. Kawanami et al. [97]
variance. The predicted values of the segregation index were shown
presented a microfluidic water-mediated reaction system based on
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 687

Table 3
Summery of micromixer for microreactor applications.

Ref. and First Year Micromixer type Flow rate Index (%) Objective Materials
author

[110] Choe 2008 SAR mixer (passive) 5–30 ␮L/min M: 97% Lithium–halogen n-BuLi, BN, NH4 Cl,
exchange reaction anthraquinone
[111] Ukita 2008 Filter mixer (passive) 5 or 10 kPa N/R Competition ELISA NP, enzyme- labeled
AP, microbeads
[112] Fang 2009 SAR mixer (passive) R = 0.01 M: 100% Boost fluid reaction C-PC, R-PE
[113] Ziegenba 2010 Grooved mixer Re = 2.2–9.5 M: 95% Liquid-sided mass CO2 , OH−
(passive) transport
[114] Hsu 2010 SAR mixer (passive) 3 mm/s M: 100% Au nanoparticles Au-NP, DNA
decorated with
oligonucleotide
[115] Ko 2011 Pillar and nozzle mixer Re = 66.5 M: 87.7% DNA ligation DH5a competent cells,
(passive) Hind III
[116] Yasukaw 2011 T-shaped mixer 1 mL/min Y: <60% Ethyl pyruvate Ethyl lactate,
(passive) production acetonitorile, VOCl3 , O2
[117] Chen 2011 SAR mixer (passive) Re = 1 M: >95% Chaos and FRET FAM, TAMRA
reaction
[118] Li 2012 Zigzag mixer (passive) Re = 248.3 M: 90% CL reaction CL, HRP
[119] Al 2012 Grooved mixer 1.68 to 6.72 mL/min M: 95% Fast exothermic and CO2 , NaOH
(passive) mass transfer reactions
[120] Guo 2013 Multi-T mixer (passive) Re = 396–9960 N/R Heat exchanger reactor BTB, Na2 HPO4
[121] Sudar 2013 Teardrop mixer Re < 100 N/R Aldol addition reactor DHA, FSA
(passive)
[122] Dong 2014 Micro-sieve mixer 10–30 mL/min M:99.9% Micro-sieve dispersion SiO2 , BaSO4
(passive) reactor
[123] Liu 2014 T-shaped mixer Re = 395–3161 M: ∼100% Micro-impinging KI, KIO3 , H3 BO3 , NaOH
(passive) stream reactors
[124] Mitic 2015 4-jet tangential mixer Re = 1000– 1500 M: >90% Mix liquids HPTS, HCl, NaOH
(passive) microreactor
[125] Dong 2015 Micro-sieve mixer 6–23 mm/s C: <10% Catalyst microreactor Cyclohexanone,
(passive) ammonia, TS-1
[126] Plouffe 2016 SZ mixer (passive) 0.8–32 mL/min M: 95% Mass transfer 4-NPA, NaOH
microreactor
[127] Watanab 2017 K-M mixer (passive) Re = 24 N/R Central collision-type ZIF-8, 2-MIM, Zn(NO3 )2
microreactor
[128] Grundtvi 2017 2- parallel streams N/R N/R Biocatalytic PEA, ACE, APH, ATA
mixer (active) microreactors

N/R: not report; C: Conversion efficiency; M: Mixing index; Y: Yield efficiency.

to be in good qualitative agreement with the experimental results have a relatively poorer mixing performance since they induced
for various agitation speeds and feed locations. a parallel flow structure and therefore resulted in a lower inter-
Hossain et al. [104] produced monodisperse poly(N-vinyl-2- phase mass transfer rate. Similar microfluidic devices for rapid
pyrrolidone) (PVP)-stabilized Pt nanoparticles (PtNPs) using an liquid–liquid reactions were also presented in [108,109]. Table 3
interdigital triangular micromixer. It was shown that the microflu- summarizes the proposals presented in [110–128] for the further
idic mixing process suppressed the hydrolytic decomposition use of micromixers in microreactor applications.
of the BH4− ions by the PtNPs formed in the initial stage of
the reaction. As a result, the PtNP yield was higher than that
obtained using a conventional batch mixing approach. Zuidhof et al. 5. Polymerization process [129–158]
[105,106] performed the Beckmann rearrangement of cyclohex-
anone oxime to ε-caprolactam in a microreactor with an integrated Mixing in polymerization processes is an extremely important
SAR micromixer. The microreactor consisted of a low-temperature issue since a non-uniform mixing inevitably degrades the proper-
mixing zone (65 ◦ C) followed by a high-temperature reaction zone ties of the final synthesized product. Micromixers [129,130] can be
(100–127 ◦ C) (Fig. 12). The specified temperature conditions were employed either for the initial mixing of the reactant streams or the
shown to achieve a selectivity of 99%. By contrast, that reported mixing of the reactive solutions encountered in multi-stage block
in the literature for a similar setup but with a uniform mixing and copolymerization processes. Moreover, micromixers with specific
reaction temperature in the range of 100–127 ◦ C was just 95%. It designs can be employed to promote gradients of functionalities
was additionally shown that the microdroplet size ranged from along the polymer backbone. In addition, when coupled with liv-
10–25 ␮m, while the residence time of the reactants was less than ing or controlled radical polymerization techniques, micromixers
40 s. Finally, the molar ratio of oleum to cyclohexanone oxime was allow for the synthesis of multi-block copolymers, whose over-
reduced to 0.8 from the industrial value of 1.2. all polydispersity indices and average molecular weights depend
Plouffe et al. [107] examined the flow regimes and mass transfer directly on the geometry and characteristic dimensions of the
rates in the two-phase alkaline hydrolysis of 4-nitrophenyl acetate microsystem [131].
given five different complex micro-reactors with different mixing Honda et al. [132] performed the continuous polymerization
mechanisms (Fig. 13). It was found that in systems with a low inter- of N-carboxy anhydrides in a microreaction system consisting
facial tension, the flow structure changed from slug, to parallel, of a multilayer laminar passive micromixer and PTFE micro-
and then to drop flow as the flow rate was increased. However, tubes (Fig. 14). The properties of the synthesized products were
in systems with a higher interfacial tension, parallel flow was not compared with those obtained in a conventional batchwise pro-
observed. In general, the curvature-based reactors were found to cess. It was shown that the polymers produced in the proposed
microreaction system had a narrower molecular weight distri-
688 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

Fig. 16. Schematic illustration of continuous flow reactor with HP-IMM micromixer for production of different end groups. (Reprinted from Ref. [137] with permission of
American Chemical Society.).

Fig. 17. (a) Sn(OTf)2 catalyzed continuous flow ring-opening polymerization in PTFE tubular microreactor with T-type micromixer, and (b) SEC of PCL with DPtarget = 10 at
80 ◦ C and DPtarget = 30 at 110 ◦ C in microreactor and batch reactor. (Reprinted from Ref. [139] with permission of Royal Society of Chemistry.).

bution than those produced in the batchwise process. Moreover, tion devices based on various types of micromixer design were
the molecular weight of the synthesized products could be con- presented for rapid polymerization applications in [135–140]. Ton-
trolled by adjusting the flow rates of the reactants injected into hauser et al. [136] synthesized end-functionalized polystyrenes
the microreactor. Rosenfeld et al. [133] performed the continuous by living anionic polymerization in a microstructured reactor via
nitroxide-mediated copolymerization of n-butyl acrylate (BA) and termination by acetal-protected functional epoxides. In the pro-
styrene (S) in a two-stage process based on two serial microtube posed method, the monomer and indicator were mixed using a
reactors. For comparison purposes, the output flow from the first high-pressure slit interdigital micromixer (HP-IMM) immersed in
reactor was mixed with the second monomer solution using two a water bath with a predefined temperature. The reaction mix-
different types of micromixer, namely a high pressure interdigi- ture was then combined with the termination reagent (glycidyl
tal multilamination micromixer (HPIMM) and a simple T-junction ether) in a simple T-junction micromixer. The end-functionalized
micromixer. The results showed that the use of the HPIMM led to polymers were recovered within several seconds. Moreover, quan-
a narrower molecular weight distribution of the synthesized prod- titative end-capping was confirmed in all cases by MALDI-ToF-MS
ucts (i.e., an improved control of the copolymerization reaction) observations. In a later study by the same group [137], a continu-
due to a more efficient mixing of the viscous n-butyl acrylate (BA) ous flow reactor with a tangential four-way jet micromixing device
and the liquid styrene prior to entering the second microreactor. was used to synthesize living polymers with a predictable disper-
Kessleret et al. [134] performed the rapid polycondensation sity and a controlled molecular weight through a suitable control
of trialkoxysilanes to PSSQs in a continuous-flow microreactor of the total flow rate (Fig. 16).
consisting of two syringe pumps, two preheating loops, a SAR Zhu et al. [139] presented a simple PTFE tubular microreactor
micromixer, and Mecanyl tubes of different lengths (1, 3, 6, 9 or with a T-type micromixer (Fig. 17(a)) for the ring opening polymer-
12 m) to adjust the residence time (Fig. 15). It was shown that ization of cyclic monomers using Stannous(II) trifluoromethane
the molecular weight of the synthesized products increased from sulfonate (Sn(OTf)2 ) as a catalyst. It was shown that under the
1900–11,000 g/mol as the residence time increased. In other words, same reaction conditions, the continuous mode reactor resulted
the microreactor setup offered the possibility to synthesize PSSQs in faster CL polymerization than that achieved in a conventional
with adjustable properties. Similar continuous flow microreac- batch mode operation (i.e., 90.1% vs. 81.8% monomer conversion
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 689

Fig. 18. (a) Photograph of suspension injection with T-shape micromixer, and (b) mixing effect of fluorescent slugs. (Reprinted from Ref. [165] with permission of Elsevier.).

Table 4
Summery of micromixer for polymerization process applications.

Ref. and First Year Micromixer type Flow rate Efficiency (%) Objective Materials
author

[138] Qiu 2016 T-junction mixer 40 mL/min C: 80% Acrylic acid polymerization K2S2O8, Acrylic acid
(passive)
[140] Nagaki 2004 IMM, SAR mixer 10 mL/min C: 80% Cation pool-initiated Vinyl Ethers, N- acrliminium
(passive) controlled polymerization ion
[141] Iwasaki 2005 T-shape mixer 0.2–1.0 mL/min Y: 34.2% Polymerization of BA, BMA, BA, BMA, MMA,VBz, St
(passive) MMA,VBz, and St
[142] Yasuhiro 2006 Y-shape mixer 2.5–5.0 mL/min C: 100% Bond forming reaction PA-PAT, (NH4 )2 PdCl4,
(passive) (PA-TAP-Pd)
[143] Iwasaki 2007 T-shape (passive) 0.17 mL/s C: 100% Cationic polymerization of CF3 SO3 H, 1,3-DPB
vinyl ethers
[144] Rosenfeld 2008 HIPMM, and slit plate 9.3–26.5 mL/min C: 98% Copolymerization of BA and St BA, St, Toluene, acetic
mixer (passive) anhydride
[145] Nagaki 2008 T-shape mixer (4, 2.0–3.0 mL/min Y: 100% Anionic polymerization THF, Me2 SiCl2 , s-BuLi
passive)
[146] Nagaki 2009 T-shape mixer (2, 3.0 mL/min C: 100% Anionic polymerization MMA, But MA, BuMa, s-BuLi
passive)
[147] Iida 2009 T-mixer (active) 0.5–10 mL/h C: 97% Anionic polymerization s-BuLi, NaOH, CaH2
[148] Nagaki 2010 T-shape mixer (3, 3.12–3.25 mL/min C: 99% Anionic polymerization MMA, But MA, BuMa, s-BuLi
passive)
[149] Nagaki 2011 T-shape mixer (2, 1.0–3.0 mL/min C: 100% Anionic polymerization MMA, But MA, BuMa, MeOH
passive)
[150] Nagaki 2012 T-shape mixer (4, 1.75–7.0 mL/min C: 100% Anionic polymerization LiCl, THF, alkyl methacrylates,
passive) tert-butyl acrylate
[151] Venden 2013 SOR-2 static mixer 1–3 mL/min C: 80% RATF polymerization THF, PnBuA, CPD-TTC, DoPAT
(passive)
[152] Méndez 2014 SAR and LLMR mixer 420 mL/min C: 30% Free-radical polymerization L-A75, L-331M80, L-JWEB50,
(passive) MEK
[153] Haven 2015 SOR-2 static mixer 2–5 mL/min Y: 14, 43% RATF polymerization nBA, EHA, AIBN, CPD-TTC,
(passive) DoPAT
[154] Zhang 2015 Zigzag-shape mixer 2.4 kPa N/R Interfacial polymerization TMC, POSS, PIP, JEFF
(passive)
[155] Nagaki 2015 T-shape mixer (3, 3–15 mL/min Y: 99% Cationic Polymerization 2,6-di-tert-butylpyri-dine,
passive) trimethyl(1-phenylv-
inyloxy)silane,
allyltrimethylsilane
[156] Tani 2016 T-shape mixer (2, 5 mL/min Y: 99% Cationic polymerization Cu-AAC
passive)
[157] Nagaki 2016 T-shape mixer (2, 10–30 mL/min Y: 100% Anionic polymerization s-BuLi, n-BuLi, THF, MeOH
passive)
[158] Ryu 2017 T-shape mixer 2, 6 mL/min N/R styrene radical polymerization Toluene, AIBN, styrene
(passive)

N/R: not report; C: Conversion efficiency; Y: Yield efficiency.

for 40 min). Furthermore, the continuous mode reactor suppressed 6. Extraction and purification processes [159–191]
inter/intramolecular transesterification due to its high mixing and
mass/heat transfer efficiency, and therefore resulted in a narrower Extraction using microfluidic devices is generally applied to
molecular weight distribution (Fig. 17(b)). Table 4 summarizes the solvent extraction, also known as liquid–liquid extraction. The
applications of micromixers for polymerization processes reported extraction process consists of two steps, namely the formation
in [138–158]. of a dispersion followed by phase separation. The main parame-
ter governing the extraction process is mass transport across the
690 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

organic phase tended to form a cap around the rear end of the
aqueous slug. Malsche et al. [166] investigated the liquid–liquid
extraction performance of three different mesoflow reactors. For
each mixer, the mass transfer rate and extraction performance were
evaluated using a fluorescence observation method. The LTF-MX
mixer was found to provide the highest extraction performance
at high flow rates, while the vortex-reactor was found to offer a
superior performance at intermediate flow rates. In a later study
[167], the same group combined a continuous flow reactor with
a self-built settler to perform the separation of Co/Ni with cyanex
272. It was shown that the use of the settler resulted in a signif-
icant reduction in the total residence time in both a single-stage
setup and a multi-stage setup. Furthermore, it was shown that the
useful volume of the proposed system was much higher than that
of a conventional mixer-settler system (i.e., a phase-separation-
volume-to equilibrium-volume ratio of 1:2 as opposed to 4:8 for a
conventional system).
The quality of processes such as polymerase chain reaction (PCR)
is fundamentally dependent on the sample condition. Thus, to
improve the detection accuracy of pathogens (for example), purifi-
Fig. 19. Schematic illustration of active microfluidic T-mixer for DNA digestion and
cation of the DNA sample is first required before the PCR process can
restriction. (Reprinted from Ref. [171] with permission of AIP Publishing LLC.).
be performed. Furthermore, a higher sample concentration is also
beneficial in improving the detection sensitivity. Consequently,
various micromixers based on a filtering or trapping concept have
phase boundary. Consequently, efficient micromixers are essential been proposed for controlling the buffer concentration or generat-
in improving the extraction performance [159–162]. ing chaotic advection [168–170].
Mae et al. [163] presented two SAR micromixers (YM-1 and Lin et al. [171,172] presented an integrated microfluidic chip for
YM-2) for the rapid aqueous extraction of phenol from dodecane. rapid DNA digestion and time-resolved capillary electrophoresis
It was shown that both micromixers enabled the production of (CE) analysis. The chip comprised two gel-filled chambers for DNA
soap-free emulsions after very short contact times. In addition, enrichment and purification, respectively, an active microfluidic
the emulsion droplet size reduced with an increasing total flow T-mixer for DNA/restriction enzyme mixing (Fig. 19), a serpen-
rate or decreasing oil concentration. The two micromixers achieved tine channel for DNA digestion reaction, and a CE channel for
maximum emulsion production rates of 5 L/h (YM-1) and 20 L/h on-line CE analysis. The time-resolved electropherograms showed
(YM-2), respectively. It was therefore concluded that both devices that the device was capable of concentrating and analyzing Фx-
had the potential for application in large-scale industrial plants. 174 DNA samples comprising 11 fragments within 24 min. Similar
Sprogies et al. [164] evaluated the performance of seven differ- DNA digestion microfluidic devices based on three different types
ent micromixers when applied to the extraction of iodine from an of passive micromixer (namely, zigzag, spiral and split-and-merge
aqueous solution into n-hexane. The results showed that a mul- (SAM)) were proposed in [173–175].
tilamination mixer achieved a better extraction efficiency than a Lee et al. [176] presented a serpentine chaotic micromixer for
T-junction mixer (i.e., 0.75 and 0.5, respectively). However, the DNA purification purposes, in which fluid twisting and flattening
extraction efficiency was still significantly lower than that achieved mixing effects were induced in two channels arranged perpendic-
by a nozzle mixer and SAR mixer (i.e., 0.9 and 0.95, respectively). ularly to one another. The effects of the microchannel geometry
Ufer et al. [165] investigated the potential for using suspended on the mixing performance were systematically explored using
catalyst particles in two-phase slug flow to perform the hydro- two Tris-HCl buffer solutions with and without fluorescein label-
genation of m-nitrotoluene with aqueous potassium formate in a ing, respectively. Moreover, the DNA purification performance was
T-shaped micromixer (Fig. 18). The particle behavior was observed confirmed using a sample obtained by dissolving a single hair
in various biphasic liquid–liquid systems using a fluorescence illu- root in DNA extraction solution. The results showed that the pro-
mination technique. In general, the results showed that silicon posed micromixer greatly simplified the DNA preparation process
dioxide or aluminum oxide particles suspended in the aqueous compared to traditional methods. Lee and Voldman [177] investi-
phase followed the flow streamlines within the mixer almost gated the performance of three different micromixer designs (i.e.,
exactly, while carbon-based particles suspended in continuous

Fig. 20. (a) Schematic illustration of assembled device, (b) particle trapping in laminar flow and chaotic flow, and (c) three different micromixer geometries. (Reprinted from
Ref. [177] with permission of American Chemical Society.).
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 691

Fig. 21. Schematic illustrations of single-chamber micromixing and PCR. (Reprinted from Ref. [205] with permission of American Chemical Society.).

Fig. 22. (a) Schematic view of lab-on-a-disc concept, (b) microfluidic chambers for loading LAMP reagent, mixing, metering, sealing, isothermal amplification, detection and
temperature sensing, (c) Salmonella detection processing steps. (Reprinted from Ref. [209] with permission of Elsevier.).

a groove micromixer (SGM), a herringbone micromixer (HM), and 7. Biological analysis processes [192–231]
a staggered herringbone micromixer (SHM)) in enhancing the par-
ticle trapping efficiency of DEP-based microconcentrators (Fig. 20). Micromixers have been employed in many biological analyti-
The results showed that chaotic mixing helped expose more par- cal processes, including stem cell differentiation, cell culturing, cell
ticles to the DEP field than non-chaotic mixing, and therefore lysis, immune response analysis, cancer metastasis, and DNA analy-
enhanced the trapping efficiency. Overall, the SHM mixer was sis [192–203]. Micromixers are commonly incorporated within LOC
found to provide the optimal performance; with a trapping effi- devices, which integrate several functional miniaturized devices
ciency around 50% higher than that obtained using a straight and (including micropumps, microvalves and micro-reaction cham-
smooth channel. Table 5 summarizes the proposals presented in bers) to accomplish the complete assay of bio-materials on a single
[160–191] for the use of micromixers in extraction and purification microfluidic platform. Micromixers also play a key role in PCR or
processes. loop-mediated isothermal amplification (LAMP) devices in the bio-
logical and biochemical engineering fields [204–212].
692 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

Table 5
Summery of micromixer for extraction and purification process applications.

Ref. and First Year Micromixer type Flow rate Index (%) Objective Materials
author

[160] Gürsel 2016 Inline mixer (passive) 10–50 mL/min E: 100% Liquid–liquid Toluene–water–
extraction acetone, acetate–
water–acetone
[161] Azimi 2017 T-type mixer (passive) Re = 75 E: 88.4% Liquid-liquid mass Fe3 O4
transfer
[172] Fu 2007 Doulbe-T mixer 150 V/cm, 5 Hz M: 98% DNA digestion -DNA, Eco RI enzyme
(active)
[173] Kefala 2015 Zigzag, aspiral, SAM Re = 0.5 M: 92% DNA digestion Endonuclease, 635-bp
mixer (passive) DNA
[174] Kastania 2016 Zigzag mixer (passive) 1, 10 mL/min R: 96% DNA purification Salmonella cell, DNA
M: 95%
[175] Tserepi 2014 Zigzag mixer (passive) 3 ␮L/min M: 100% DNA digestion 273-bp DNA, enzyme
Re = 0.208
[178] Kuo 2015 fishbone filtration 0.02 ␮L/s N/R Extraction of human Whole blood, plasma
mixer (passive) plasma
[179] Azimi 2015 Y-type mixer (passive) 5–20 ␮L/min E: >45% Liquid–liquid mass Fe3 O4 , n-butanol
transfer
T: >70%
[180] Domíng 2016 Multilamination, 30 mL/h N/R Supercritical fluid Ethanol, scCO2
T-type mixer (passive) extraction
[181] Shih 2016 Multi-channel mixer 400 ␮L/min E: 100% R: 83–110% Solid phase extraction Mn, Co, Ni, Cu, Pb
(passive)
[182] Jafari 2016 Twisted mixer Re = 77–460 E: 60–100% Liquid–liquid Cu(II),D2EHPA
(passive) extraction
[183] Kakavan 2016 T mixer with PJ, PJC, Re = 181 E: 77.8% (PMN) Liquid-liquid mass 1-octanol, propionic
PJN, PMC, PMN transfer acid
(passive)
[184] Rahimi 2017 CPJ, Y mixer (passive) 0.75–6 ␮L/min R0 : >95% Liquid-liquid mass Fe3 O4 , Cu (II), D2EHPA
transfer
[185] Peng 2017 Y-type mixer (passive) 20–80 ␮L/min E: ∼80% Solid phase extraction Cr(III),Cr(VI)
[186] Domíng- 2017 Multilamination, 25 ␮L/min R: >85% Supercritical fluid Ethanol, scCO2
uez T-type mixer (passive) extraction
[187] Ji 2007 Planar spiral mixer Re < 1 N/R DNA purification DNA from human
(passive) blood
[188] Morales 2010 Vortex mixer (passive) 0.4–2 ␮L/min R: >92% Phenol extraction BSA, 0.1–10-kb DNA
[189] Tai 2011 Membrane- type mixer 0.34 ␮L/s M: 96.7% Sample pretreatment ␤-actin gene, cDNA,
(active) and hybridization mRNA
[190] Chen 2013 Groove-shape mixer 1.67–8.36 ␮L/min A: 100% Sample amplification -DNA, 99-bp DNA
(passive)
[191] Hellé 2015 Thermomixer (active) 0.3–1.0 mL/h E: 90.7% Liquid–liquid Uranium(VI), Aliquat
extraction

N/R: not report; A: Amplification efficiency; E: Extraction efficiency; M: Mixing index.


R: Recovery efficiency; R0 : removal efficiency; T: Mass transfer efficiency.

Kim et al. [205] presented a simple approach for perform- consisting of pneumatic micropumps, an active vortex-type
ing micromixing and biochemical reactions on a single platform micromixer, a pneumatic micro-injector and several microvalves
without the need for pumps by using natural convection to mix for performing the entire C-reactive protein (CRP) detection pro-
the species and capillarity regulation to achieve sample transport tocol [214,215]. The micromixer performed the mixing of the
(Fig. 21). The feasibility of the proposed device was demon- reagents and biosamples using a driving frequency of 8 Hz and an
strated by performing the PCR amplification of an influenza viral applied pressure of 20 psi. The results showed that the integrated
DNA fragment. The proposed microfluidic system was reported microfluidic system was capable of automatically completing the
to have several important advantages over conventional PCR sys- entire protocol within 30 min with a detection limit of 0.0125 mg/L.
tems, including a faster thermal response, a smaller number of Furthermore, it was shown that the CRP concentrations obtained
microfluidic device components, and a pumpless operation, which for clinical samples of human serum were in good agreement with
resulted in a shorter idle time, a simpler and cheaper fabrication those obtained using a conventional benchtop system.
process, and a manual operation. Sayad et al. [209] presented a Circulating tumor cells (CTCs) in the blood of cancer patients is
centrifugal microfluidic CD-based device for the LAMP detection an important indicator of disease progression and survival. How-
of foodborne Salmonella. The CD platform was designed to per- ever, due to the rarity of CTCs in the bloodstream, they must first
form all of the steps required to complete the pathogen detection be enriched before molecular and cellular analyses can be per-
process, including LAMP reagent loading, mixing, metering, seal- formed for clinical purposes [203]. Many integrated micromixers
ing, isothermal amplification, heating and detection (Fig. 22). The have been proposed to facilitate the perturbation, lysis, separa-
experimental results showed that the device had a detection limit of tion, isolation and detection of cells [216–221]. For example, Lin
5 × 10–3 ng/␮L when tested on tomatoes spiked with Salmonella. et al. [216] presented an integrated microfluidic device in which the
Moreover, the entire process from sample preparation to detection CTSs were continuously coated with anti-epithelial cell adhesion
was completed within 70 min in a fully automated manner. molecule-conjugated beads using a Taylor–Gortler vortex microflu-
The detection and quantification of proteins is indispensable idic mixer (Fig. 24). The feasibility of the labeling system was
in fundamental research and clinical diagnosis [213]. Lee et al. demonstrated using two different breast cancer cells. The coat-
[214,215] presented an integrated microfluidic chip (Fig. 23) ing efficiency achieved using 3-mm beads was found to be very
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 693

Fig. 23. (a) Photograph showing detailed components of integrated microfluidic chip (Reprinted from Ref. [214] with permission of Elsevier.), and (b) schematic illustration
of active vortex-type micromixer. (Reprinted from Ref. [215] with permission of Elsevier.).

Fig. 24. (a) Schematic illustration showing size amplification of cells on chip, (b) microscopic image of silicon mold used to fabricate micromixer microchannel, and (c)
photograph of Taylor–Gortler vortex microchip. Note that all of the channels have the same mixing zone, composed of 20 consecutive wavy ducts. (Reprinted from Ref. [216]
with permission of Elsevier.).

different in each case. Thus, it was inferred that the labeling sys- experimental results showed that the proposed device achieved a
tem had good antigen-antibody specificity. Moreover, the device detection time of just 30 min (around 8 times faster than a con-
showed a high-throughput performance with a treatment time ventional benchtop method). Moreover, the detection limit was
of just 11.25 min for a 7.5 mL blood sample and a flow rate of found to be 21 pg (around 38 times better than that of a traditional
600 mL/min. method). Zhang et al. [227] developed a microfluidic mixer-based
Bioterrorism and bio-warfare have emerged as major concerns method for determining the inactivation kinetics of Escherichia coli
in recent years. Cho et al. [222] developed an integrated microflu- in chlorine solutions. The mixer incorporated three different mixing
idic device for performing the bio-barcode assay-based detection of designs, namely Y-junction, Dean vortex, and chaotic, to accurately
biological agents, including botulinum toxin A. The device incorpo- determine the reaction time and control the chlorine concentration
rated micropumping actuation for sample loading, passive mixing (Fig. 26). It was shown that the device enabled the pathogen inac-
to induce MM–biological agent–PM complexes [223], magnetic tivation kinetics to be assessed in less than 1 s. Table 6 presents a
separation to purify the target–probe complexes, and ␮CE to ana- brief review of the other main applications of micromixers in the
lyze the barcode DNAs (Fig. 25). The device was shown to have a biological analysis field reported in [193–231].
high sensitivity and a total response time of just 30 min.
Many microfluidic devices have been proposed for virus and 8. Droplet/emulsion and other processes [232–260]
bacteria detection and evaluation purposes [224–229]. Lee et al.
[225] presented an integrated microfluidic system based on virus- Microfluidic droplet (or emulsion) devices are designed to
bound magnetic bead complexes for the rapid diagnosis of dengue enable chemical reactions to be carried out inside microdroplets
virus. The device comprised two micropumps, a pneumatic four- with a volume on the order of a few microliters or less. Gener-
membrane-type micromixer, a microcoil array and a sample ally speaking, the microfluidic droplets or emulsions are formed
purification chamber. The device permitted the simultaneous and suspended in an immiscible phase such as oil, water in oil, or
detection of IgG and IgM in two different detection chambers. The water/oil/water. Micromixing is then achieved by means of chaotic
694 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

Fig. 25. (a) Schematic illustration of biobarcode assay based biological agent detection, (b) photograph of assembled micro device for micropumping, passive mixing,
magnetic separation, and ␮CE, and (c) photograph of assembled micro device showing sample and probe inlets, micropump and passive mixer. (Reprinted from Ref. [222]
with permission of Royal Society of Chemistry.).

Fig. 26. (a) Schematic illustration of microfluidic mixer that simulates instant cross-contamination and sanitization scenarios, and (b) photograph of sample microfluidic
device used to simulate pathogen survival scenarios. (Reprinted from Ref. [227] with permission of Elsevier.).

advection induced by either droplet motion or the channel shape and amino functional groups to form a doubly-crosslinked nanoad-
[232–245]. hesive (DCNA) layer (Fig. 27). The improved bonding strength
Yeh et al. [236] presented a gradient-microfluidic droplet gen- provided by the DCNA allowed the mixer to be operated with water
erator consisting of J-junction block passive micromixers and flow rates as high as 40 mL/min (corresponding to a Reynolds num-
flow-focusing devices to generate droplets with four different sizes ber of Re = 4423). Notably, the high flow rate enabled a yield of
and concentrations. It was shown that the size of the droplets as much as 10 mg of emulsion to be obtained within one minute.
could be precisely controlled by adjusting the relative flow rates The corresponding throughput (4.0 × 106 droplets per second) was
of the water phase and oil phase, respectively. The feasibility of the thus far higher than that obtained using conventional droplet
proposed platform was demonstrated by generating Ca-alginate microfluidic devices with a typical throughput of just a few tens
microcapsules with different concentrations of BSA for drug release of thousands.
purposes. In a later study [237], the same group developed a similar Lee et al. [243] developed a magnetic droplet microfluidic sys-
microfluidic droplet generator to produce water droplets with 11 tem in which the mixing performance was enhanced by means of
different trypan blue concentrations. The water droplets were then acoustic excitation (Fig. 28(a)). The experimental results obtained
used to encapsulate magnetic Fe3 O4 nanoparticles for potential using droplets of methylene blue and glycerol (5 wt%) showed
use in the drug delivery or pharmaceutical fields. The experimen- that a fully mixed condition could be achieved within 60 s when
tal results showed that the chitosan emulsions had a coefficient of acoustic excitation was applied, whereas the droplets remained
variation of less than 10% and a microparticle size of 44–83 ␮m. unmixed in the absence of excitation. In addition, it was shown
Turbulent flow provides an effective means of achieving homo- that droplets of a different size were excited at different acous-
geneous mixing. However, generating turbulent flow in microscale tic frequencies. As a result, the device offered the possibility for a
channels presents a major challenge. Accordingly, You et al. [242] selective mixing performance. Samiei et al. [246] presented an elec-
fabricated a Y-shaped turbulence microfluidic mixer in which the trohydrodynamic (EHD) technique for performing the rapid mixing
bonding strength between the PDMS and glass substrates was of stationary droplets on digital microfluidic platforms by apply-
enhanced via a chemical surface treatment process using epoxy ing a high frequency AC voltage to induce circulation zones within
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 695

Fig. 27. (a) Experimental setup for measuring pressure drop along microfluidic channel, and (b) intensity profiles, corresponding snapshot images, and simulated concen-
tration contour images at inlet and outlet. (Reprinted from Ref. [242] with permission of Royal Society of Chemistry.).

Table 6
Summery of micromixer for biological analytical process applications.

Ref. and First Year Micromixer type Flow rate Index (%) Objective Materials
author

[193] Chen 2007 Filtration mixer 1–25 ␮L/min M: ∼90% Cell separation, lysis and DNA Whole blood.
(passive) purification
[194] Huh 2007 Actuated mixer (active) 1.96 Hz M: ∼90% Cell disruption Escherichia coli
[196] Nason 2011 Zig-zag shaped mixer 0.3–3 ␮L/min M: 79.4% Drug screening SaOS-2 cell
(passive)
[197] Wang 2012 Ribcage mixer 0.05 mL/h M: ∼100% Protein concentration Bovine serum albumin (BSA)
(passive) measurements
[198] Balbino 2013 Vortex mixer (passive) 140 mm/s N/R Produce pDNA/CL complexes. Plasmid DNA, cationic
liposome
[199] Lin 2014 Bubble-driven mixer 25–400 mL/h M: 90% Bladder cancer biomarker Bladder cancer
(active) detection
[200] Aguirre 2015 Trapezoid type mixer Re = 11 B:2% v/v blood Incubation and separation MCF7 breast cancer cell
(passive) cancer cell
[203] Lee 2017 Multivortex mixer 400 ␮L/min R: 90.63% Isolation of circulating tumor MCF-7 cell
(passive) cells
M: ∼80%
D: ∼99%
[206] Choi 2012 Multi-layer mixer 12 ␮L/min N/R PCR-microarray system 11 types genes
(passive) (100-bp ∼ 290-bp)
[208] Geissler 2014 Air stream mixer Re ∼100 M: ∼90% DNA hybridization assays L. monocytogenes, P. ramorum
(passive) and V. dahlia DNA
[211] Louns- 2013 Serpentine mixer 10 ␮L/min A: 6-fold PCR system ␤-globin, gelsolin genes
bury (passive)
[213] Li 2013 Dual-focusing mixer 35 kPa M: 93% DNA–protein interaction G-quadruplex, SSBP
(passive)
[215] Yang 2015 Vortex-type mixer 300 ␮L/min M: > 96% (7 Hz) C-reactive protein C-reactive protein
(active) measurement
[217] Rajabi 2014 SAR and SHM mixer 2000 ␮L/min CL: 100% Cell perturbation, lysis, and CHO cell
(passive) separation
[218] Femmer 2015 SHM mixer (passive) Re = 10 N/R Gas-liquid contact oxygenation Red blood cell
device
[220] Cosen- 2015 3-SERP mixer (passive) Re = 0.1 M: 93% Red blood cell lysis Red blood cell
tino
[223] Gao 2015 Acoustic mixer 125–150 Hz B: >50% Antibody-antigen binding Antibody-antigen
(passive) assay
[226] Wang 2013 USCC mixer (passive) Re > 80 M: >80% Bacteria detection and Escherichia coli
quantification
[229] Petkovic 2017 Rotary mixer (active) Re = 0.02 N/R Detection of Hendra virus Hendra virus
[230] Jung 2011 Serpentine 3D mixer 3.8–100 ␮L/h CP: 100% Multiplex pathogen detection 3 bacteria cells, barcode DNA
(passive)
[231] Liu 2016 3D U-type mixer 0.21 ␮L/min M: >90% Analysis of biomacromolecules G-quadruplex
(passive) folding kinetics

N/R: not report; A: Amplification efficiency; B: Binding efficiency; CL: Cell lysis efficiency.
CP: Cell capture efficiency; D: Depletion efficiency; M: Mixing index; R: Recovery efficiency.

the droplets, thereby prompting a rapid mixing of their contents In addition to the critical roles played by micromixers in the
(Fig. 28(b)). It was reported that the proposed platform had sev- chemical and biological applications described above, micromixers
eral key advantages over conventional mixing methods, including can also be used to perform important functions such as microflu-
a shorter mixing time, a lower number of electrodes (and hence a idic switching; nanocomplex, oxidation and adduct formation;
smaller size), and a stationary operation (i.e., a reduced risk of cross velocity measurement; supercritical fluid fractionation; and so on
contamination and surface bio-fouling). [247–253]. Bezagu et al. [248] presented a rapid active mixer based
696 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

Fig. 28. (a) Schematic illustrations of magnetic droplet manipulation system actuated by magnetic actuation and acoustic excitation (Reprinted from Ref. [243] with permis-
sion of Elsevier.), and (b) schematic illustrations of electrohydrodynamic mixing in open DMF and closed DMF systems. (Reprinted from Ref. [246] with permission of Royal
Society of Chemistry.).

Fig. 29. Ultrasound-induced vaporization of a perfluorocarbon phase for efficient mixing in microfluidic channels. (Reprinted from Ref. [248] with permission of Royal Society
of Chemistry.).

on the localized vaporization of a perfluorocarbon stream sand- reagents within seconds after they were generated and with an
wiched between two fluid streams at the focal zone of an ultrasound efficiency of ∼80% (i.e., ∼8-fold higher than that achieved in a typ-
transducer (Fig. 29). It was shown that a thorough mixing of the ical T-junction mixer). Table 7 presents a brief review of the other
PFC stream with the adjacent fluid streams was achieved within main proposals for the use of micromixers for droplet/emulsion and
approximately 100 ms after the application of an acoustic pulse. other processes reported in [234–260].
Moreover, laminar flow was re-established within approximately
the same time scale; thereby enabling the PFC phase and mixed
phase to be easily separated at the outlets of the mixing channel. 9. Conclusions
Van den Brink et al. [251] developed a microfluidic electrochem-
ical cell with an integrated passive gradient rotation micromixer for Micromixers continue to receive significant attention in the
the oxidation and adduct formation of xenobiotics. The micromixer research community due to the tremendous advantages they bring
was designed specifically to mix liquids in the shallow channels to the fields of chemical, biological, medical and environmental
required for thin-layer electrochemical flow cells. Moreover, its analysis. Since the year 2000, more than 2300 papers focusing
operating principle was based on a rotation of the concentration on the topic of micromixers have been published. Furthermore,
gradient over the entire channel height. The gradient rotation mixer these papers have been the subject of more than 40000 citations
enabled the reactive oxidation products to be mixed with additional (January 2010–June 2017: ISI Web of Science). Of these papers,
around 20%–25% relate to the applications of micromixing tech-
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 697

Table 7
Summery of micromixer for droplet/emulsion and other process applications.

Ref. and First author Year Micromixer type Flow rate Index or size Objective Materials

[234] Furtaw 2009 Serpentine mixer 1–3 ␮L/min Ei: 35-fold Liquid–liquid droplet Ag colloids, PBS solution
(passive) process
[235] Matsu- yama 2010 Orifice-shaped mixer 3–45 L/h D: 5–30 ␮m Emulsification process Cyclic dimethyl silicon,
(passive) aqueous solution
[238] Ribeiro- Costa 2009 LTCC mixer (passive) 1.1–2.2 L/h Ee: 61–75% Emulsion/solvent BSA, PLGA
evaporation process
[239] Matsu- yama 2011 K-JET mixer (passive) 3–50 L/h D:10–100 ␮m Emulsification process Cyclic silicone, PVA, gel
capsule
[240] Montillet 2013 T-shape mixer Re: 5–23(O), D: <10 ␮m Emulsification process Sunflower oil,
(passive) 30–163(W) demineralized water
[241] Maillot 2014 Serpentine mixer 33–165 mm/s N/R Microfluidic droplet BSA, PVB
(passive) process
[244] Chen 2017 Laminar mixer Po = 500 mbar M: 98% Microfluidic droplet AcPHF6, ThT, orange G
(passive) process
Pw = 440 mbar D: ∼100 ␮m
[245] Larrea 2017 Slit-interdigital mixer 36–54 mL/min Ee: 100% Emulsification process SiO2 , Au, PLGA
(passive)
[247] Abadi 2013 SAR mixer (passive) 40 ␮L/min M: 85% Measured pH HCl, NaOH
[249] Bohr 2017 3D circular mixer 36 mL/min PDI: 0.3 Nanoconplex PDDA, PSS
(passive) production
[250] Santillo 2012 Turbulent vortex mixer 20 mL/min Mean size 60 nm Nanoparticle MoS2
(passive) precipitation
[252] Ergin 2015 Magnetic mixer 0–70 mm/s N/R Micro PIV Nile-red fluorescent,
(active) ␥-Fe2 O3
[254] Guo 2011 Serpentine mixer 30–80 ␮L/h D: 1–1.3 nL Microfluidic droplet Soybean oil, HgCl2 , Hg(II)
(passive) process
[255] Giuffrida 2015 Serpentine mixer 0.2–1 ␮L/min D: 20 nL Microfluidic droplet- K562 cell, miR-201, dNTPs,
(passive) ICSDP process RNaseOUT
[256] Ochs 2015 Serpentine mixer 0–67 psi N/R DNA synthesis in GGA compatible DNA,
(passive) droplet process
[257] Jin 2015 Bumpy mixer (passive) 100–500 ␮L/h D: ∼70 ␮m Microfluidic droplet Mineral oil, PBS, cell,
process
[258] Qu 2017 Heating mixer (active) 0.5 W, 1.56 GHz M: ∼100% Microfluidic droplet DMOS, xylene
process
Y: 60%
D: 20 ␮L
[259] Akbari 2017 3 spiral mixers 50–3000 ␮L/h D: 10–60 ␮m Cell micro- PEG-4-NH2 , PEG-4-NHS,
(passive) 3.1 MHz encapsulation process MDA-MB231 cell
[260] Bokharaei 2017 Laminar jet mixer 0.05–1 ␮L/min Ee: 96% Protein drug delivery BSA, PLLA
(passive) process
D: 20–30 ␮m

N/R: not report; Ee: Encapsulation efficiency; Ei: Enhancement intensity; D: Droplet diameter (size).
M: Mixing index; PDI: Polydispersity index; R: Production Recovered; Y: Yield efficiency.

nology, while the remainder concern the fundamental element


development or enhanced micromixing performance of micromix-
ers. Fig. 30 presents a simple statistical breakdown of the published
research (from 2004 to 2017, but mostly in the past ten years)
covered in this review in terms of their related applications. As
shown, the main application fields include chemical reactors (18%),
biological analysis (17%) and chemical synthesis (16%).
Recent years have witnessed a persistent trend toward the
development of ␮TAS or LOC devices with integrated micromixers.
As described earlier in this review, micromixers can be classified as
either passive or active; depending on the manner in which species
mixing is performed. Passive micromixers contain no moving parts
and require no energy input other than the pressure head used to
drive the fluid flows at a constant rate. As a result, they are eas-
ily fabricated, designed and integrated in LOC devices. By contrast,
active microfluidic mixers stir or agitate the fluid flow using some
form of external energy supply. Thus, they tend to be bulkier than
passive mixers and less easily integrated in LOC devices. Conse-
quently, of the studies reviewed in this article, around 90% consider
passive micromixers.
Overall, the results presented in this review confirm the impor-
tance of micromixing technology for applications in the chemical,
biological, medical, environmental analysis, and chemical indus-
tries. While significant advances in micromixing technology have
Fig. 30. Statistical breakdown of published literature in micromixing field by appli-
cation type. been made in recent years, it seems likely that micromixing will
698 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

continue to thrive as an active research field for years (if not [26] Y.C. Toh, T.C. Lim, D. Tai, G. Xiao, D. van Noort, H. Yu, A microfluidic 3D
decades) to come as microfluidic and micro-electro-mechanical hepatocyte chip for drug toxicity testing, Lab Chip 9 (2009) 2026–2035.
[27] X. Yan, L. Hu, Y. Li, X. Feng, B. Liu, Microfluidic slipchip-based reaction
systems (MEMS) technologies continue to evolve and progress. microarray with dual concentration gradient, Chin. J. Anal. Chem. 43 (2015)
1520–1525.
[28] M.J. Schertzer, R. Ben Mrad, P.E. Sullivan, Automated detection of particle
Acknowledgements concentration and chemical reactions in EWOD devices, Sens. Actuators B:
Chem. 164 (2012) 1–6.
The authors would like to thank the Ministry of Science and [29] T.F. Hong, W.J. Ju, M.C. Wu, C.H. Tai, C.H. Tsai, L.M. Fu, Rapid prototyping of
PMMA microfluidic chips utilizing a CO2 laser, Microfluid. Nanofluid. 9
Technology of Taiwan for the financial support of this study (2010) 1125–1133.
under Grant Nos. MOST 103-2320-B- 020-001- MY3, MOST 103- [30] H.H. Hou, Y.N. Wang, C.L. Chang, R.J. Yang, L.M. Fu, Rapid glucose
2221-E-020-025-MY3, MOST 106-2314-B-020 −002 −MY3, MOST concentration detection utilizing disposable integrated microfluidic chip,
Microfluid. Nanofluid. 11 (2011) 479–487.
106-2221-E-020 −019 −MY3, MOST 106-2622-B-020-001 −CC2, [31] Y.N. Wang, R.J. Yang, W.J. Ju, M.C. Wu, L.M. Fu, Convenient quantification of
and 106TFDA-A-103. methanol concentration detection utilizing an integrated microfluidic chip,
Biomicrofluidics 6 (2012) 034111.
[32] X. Pana, A.C. Lewis, M. Rodenas-Garcia, Microfluidic lab-on-a-chip
References derivatization for gaseous carbonyl analysis, J. Chromatogr. A 1296 (2013)
93–103.
[1] R.J. Yang, H.H. Hou, Y.N. Wang, L.M. Fu, Micro-magnetofluidics in [33] K.S. Lok, Y.C. Kwok, N.T. Nguyen, Passive micromixer for luminol-peroxide
microfluidic systems: a review, Sens. Actuators B: Chem. 224 (2016) 1–15. chemiluminescence detection, Analyst 136 (2011) 2586–2591.
[2] Q. Deng, Q. Lei, R. Shen, C. Chen, L. Zhang, The continuous kilogram-scale [34] L.M. Fu, W.J. Ju, R.J. Yang, Y.N. Wang, Rapid prototyping of glass-based
process for the synthesis of 2,4,5-trifluorobromobenzene via Gattermann microfluidic chips utilizing two-pass defocused CO2 laser beam method,
reaction using microreactors, Chem. Eng. J. 313 (2017) 1577–1582. Microfluid. Nanofluid. 14 (2013) 479–487.
[3] R.J. Yang, C.C. Liu, Y.N. Wang, H.H. Hou, L.M. Fu, A comprehensive review of [35] P.H. Chiu, C.C. Chang, R.J. Yang, Electrokinetic micromixing of charged and
micro-distillation method, Chem. Eng. J. 313 (2017) 1509–1520. non-charged samples near nano–microchannel junction, Microfluid.
[4] C.C. Liu, Y.N. Wang, L.M. Fu, Micro-distillation system for formaldehyde Nanofluid. 14 (2013) 839–844.
concentration detection, Chem. Eng. J. 304 (2016) 419–425. [36] M. La, S.J. Park, H.W. Kim, J.J. Park, K.T. Ahn, S.M. Ryew, D.S. Kim, A centrifugal
[5] K. Meller, M. Szumski, B. Buszewski, Microfluidic reactors with immobilized force-based serpentine micromixer (CSM) on a plastic lab-on-a-disk for
enzymes-Characterization dividing, perspectives, Sens. Actuators B: Chem. biochemical assays, Microfluid. Nanofluid. 15 (2013) 87–98.
244 (2017) 84–106. [37] M. Javanmard, S. Emaminejad, C. Gupta, J. Provine, R.W. Davis, R.T. Howe,
[6] S.A. Nabavi, G.T. Vladisavljević, V. Manović, Mechanisms and control of Depletion of cells and abundant proteins from biological samples
single-step microfluidic generation of multi-core double emulsion droplets, byenhanced dielectrophoresis, Sens. Actuators B: Chem. 193 (2014)
Chem. Eng. J. 322 (2017) 140–148. 918–924.
[7] T. Kanai, M. Tsuchiya, Microfluidic devices fabricated using stereo [38] L.M. Fu, Y.N. Wang, C.C. Liu, An integrated microfluidic chip for
lithography for preparation of monodisperse double emulsions, Chem. Eng. formaldehyde analysis in Chinese herbs, Chem. Eng. J. 244 (2014) 422–428.
J. 290 (2016) 400–404. [39] L.M. Fu, W.J. Ju, C.C. Liu, R.J. Yang, Y.N. Wang, Integrated microfluidic array
[8] C.C. Liu, Y.N. Wang, L.M. Fu, D.Y. Yang, Rapid integrated microfluidic chip and LED photometer system for sulfur dioxide and methanol
paper-based system for sulfur dioxide detection, Chem. Eng. J. 316 (2017) concentration detection, Chem. Eng. J. 243 (2014) 421–427.
790–796. [40] W.M. Abed, R.D. Whalley, D.J.C. Dennis, R.J. Poole, Experimental
[9] C.Y. Lee, W.T. Wang, C.C. Liu, L.M. Fu, Passive mixers in microfluidic systems: investigation of the impact of elastic turbulence on heat transfer in a
a review, Chem. Eng. J. 288 (2016) 146–160. serpentine channel, J. Non-Newtonian Fluid Mech. 231 (2016) 68–78.
[10] N.T. Nguyen, Z. Wu, Micromixers –a review, J. Micromech. Microeng. 15 [41] C.F.A. Floquet, V.J. Sieben, B.A. MacKay, F. Mostowfi, Determination of boron
(2005) R1–R16. concentration in oil field water with a microfluidic ion exchange resin
[11] V. Hessel, H. Löwe, F. Schönfeld, Micromixers –a review on passive and instrument, Talanta 154 (2016) 304–311.
active mixing principles, Chem. Eng. Sci. 60 (2005) 2479–2501. [42] B. Ibarlucea, X. Munoz-Berbel, P. Ortiz, S. Büttgenbach, C.
[12] C.C. Chang, R.J. Yang, Electrokinetic mixing in microfluidic systems, Fernández-Sáncheza, A. Llobera, Self-validating lab-on-a-chip for
Microfluid. Nanofluid. 3 (2007) 501–525. monitoring enzyme-catalyzed biological reactions, Sens. Actuators B: Chem.
[13] C.Y. Wen, C.P. Yeh, C.H. Tsai, L.M. Fu, Rapid magnetic microfluidic mixer 237 (2016) 16–23.
utilizing AC electromagnetic field, Electrophoresis 30 (2009) 4179–4186. [43] A. Gonzalez, L. Estala, M. Gaines, F.A. Gomez, Mixed thread/paper-based
[14] C.Y. Wen, K.P. Liang, H. Chen, L.M. Fu, Numerical analysis of a rapid magnetic microfluidic chips as a platform for glucose assays, Electrophoresis 37
microfluidic mixer, Electrophoresis 32 (2011) 3268–3276. (2016) 1685–1690.
[15] Z. Li, S. Kim, Pulsatile micromixing using water-head-driven microfluidic [44] N.V. Anh, H.V. Trung, B.Q. Tien, N.H. Binh, C.H. Ha, N.L. Huy, N.T. Loc, V.T.
oscillators, Chem. Eng. J. 313 (2017) 1364–1369. Thu, T.D. Lam, Development of a PMMA electrochemical microfluidic device
[16] C.H. Yeh, C.H. Chen, Y.C. Lin, Use of a gradient-generating microfluidic for carcinoembryonic antigen detection, J. Electron. Mater. 45 (2016)
device to rapidly determine a suitable glucose concentration for cell 2455–2462.
viability test, Microfluid. Nanofluid. 10 (2011) 1011–1018. [45] B. Wang, Y. Jia, Q. Lin, A microfabrication-based approach to quantitative
[17] Y. Awwad, T. Geng, A.S. Baldwin, C. Lu, Observing single cell NF-␬B dynamics isothermal titration calorimetry, Biosens. Bioelectron. 78 (2016) 438–446.
under stimulant concentration gradient, Anal. Chem. 84 (2012) 1224–1228. [46] J.N. Kuo, H.S. Liao, X.M. Li, Design optimization of capillary-driven
[18] X. Jiang, Q. Xu, S.K.W. Dertinger, T. Fu, G.M. Whitesides, A general method micromixer with square-wave microchannel for blood plasma mixing,
for patterning gradients of biomolecules on surfaces using microfluidic Microsyst. Technol. 23 (2017) 721–730.
networks, Anal. Chem. 77 (2005) 2338–2347. [47] J.N. Kuo, Y.S. Li, Centrifuge-based micromixer with three-dimensional
[19] K. Anwar, T. Han, S. Yu, S.M. Kim, Integrated micro/nano-fluidic system for square-wave microchannel for blood plasma mixing, Microsyst. Technol. 23
mixing and preconcentration of dissolved proteins, Microchim. Acta 173 (2017) 2343–2354.
(2011) 331–335. [48] T. Han, L. Zhang, H. Xu, J. Xuan, Factory-on-chip. Modularised microfluidic
[20] S.M. Kim, M.A. Burns, E.F. Hasselbrink, Electrokinetic protein reactors for continuous mass production of functional materials, Chem. Eng.
preconcentration using a simple glass/poly(dimethylsiloxane) microfluidic J. 326 (2017) 765–773.
chip, Anal. Chem. 78 (2006) 4779–4785. [49] L. Wang, S. Ma, B. Yang, W. Cao, Xi. Han, Morphology-controlled synthesis of
[21] J. Kim, D. Taylor, N. Agrawal, H. Wang, H. Kim, A. Han, K. Rege, A. Jayaraman, Ag nanoparticle decorated poly(o-phenylenediamine) using microfluidics
A programmable microfluidic cell array for combinatorial drug screening, and its application for hydrogen peroxide detection, Chem. Eng. J. 268
Lab Chip 12 (2012) 1813–1822. (2015) 102–108.
[22] D. Friedrich, C.P. Please, T. Melvin, Design of novel microfluidic [50] J. Dai, X. Yang, M. Hamon, L. Kong, Particle size controlled synthesis of CdS
concentration gradient generators suitable for linear and exponential nanoparticles on a microfluidic chip, Chem. Eng. J. 280 (2015) 385–390.
concentration ranges, Chem. Eng. J. 193–194 (2012) 296–303. [51] Y. Laribi, P. Perrichon, C. Berguerand, L. Kiwi-Minsker, E. Sulman, G. Szirbik,
[23] C.s Escobedo, Y. Chou, M.d Rahman, X. Duan, R. Gordon, D. Sinton, A.G. H. Richert, J.M. Lang Gottfried, S. Roggan, Chemoselective three-phase
Brolob, J. Ferreira, Quantification of ovarian cancer markers with integrated hydrogenation of an Ombrabulin nitro-stilbene intermediate in a
microfluidic concentration gradient and imaging nanohole surface plasmon continuous-flow mobile platform, Chem. Eng. J. 316 (2017) 1069–1077.
resonance, Analyst 138 (2013) 1450–1458. [52] B. Swain, M.H. Hong, L. Kang, B.S. Kim, N.H. Kim, C.G. Lee, Optimization of
[24] M. Polinkovsky, E. Gutierrez, A. Levchenko, A. Groisman, Fine temporal CdSe nanocrystals synthesis with a microfluidic reactor and development of
control of the medium gas content and acidity and on-chip generation of combinatorial synthesis process for industrial production, Chem. Eng. J. 308
series of oxygen concentrations for cell cultures, Lab Chip 9 (2009) (2017) 311–321.
1073–1084. [53] D.J. Kim, H.J. Oh, T.H. Park, J.B. Choo, S.H. Lee, An easily integrative and
[25] R.H. Lam, M.C. Kim, T. Thorsen, Culturing aerobic and anaerobic bacteria and efficient micromixer and its application to the spectroscopic detection of
mammalian cells with a microfluidic differential oxygenator, Anal. Chem. 81 glucose-catalyst reactions, Analyst 130 (2005) 293–298.
(2009) 5918–5924.
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 699

[54] J. Wang, G. Sui, V.P. Mocharla, R.J. Lin, M.E. Phelps, H.C. Kolb, H.R. Tseng, [80] M. Thiele, A. Knauer, A. Csáki, D. Mallsch, T. Henkel, J.M. Köhler, W. Fritzsche,
Integrated microfluidics for parallel screening of an in situ click chemistry High-throughput synthesis of uniform silver seed particles by a continuous
library, Angew. Chem. Int. Ed. 45 (2006) 5276–5281. microfluidic synthesis platform, Chem. Eng. Technol. 38 (2015) 1131–1137.
[55] Y. Wang, W.Y. Lin, K. Liu, R.J. Lin, M. Selke, H.C. Kolb, N. Zhang, X.Z. Zhao, [81] N. Veldurthi, S. Chandel, T. Bhave, D. Bodas, Computational fluid dynamic
M.E. Phelps, C.K.F. Shen, K.F. Faull, H.R. Tseng, An integrated microfluidic analysis of poly(dimethyl siloxane) magnetic actuator based micromixer,
device for large-scale in situ click chemistry screening, Lab Chip 9 (2009) Sens. Actuators B: Chem. 212 (2015) 419–424.
2281–2285. [82] H. Kim, K. Min, K. Inoue, D. Im, D. Kim, J. Yoshida, Submillisecond organic
[56] S.I. Kawasaki, Y. Xiuyi, K. Sue, Y. Hakuta, A. Suzuki, K. Arai, Continuous synthesis: outpacing Fries rearrangement through microfluidic rapid
supercritical hydrothermal synthesis of controlled size and highly crystalline mixing, Science 352 (2016) 691–694.
anatase TiO2 nanoparticles, J. Supercrit. Fluids 50 (2009) 276–282. [83] S. Baek, S. Song, J. Lee, J. Kim, Nanoscale diameter control of sensory
[57] S. Li, S. Meierott, J.M. Kohler, Effect of water content on growth and optical polydiacetylene nanoparticleson microfluidic chip for enhanced
properties of ZnO nanoparticles generated in binary solvent mixtures by fluorescence signal, Sens. Actuators B: Chem. 230 (2016) 623–629.
micro-continuous flow synthesis, Chem. Eng. J. 165 (2010) 958–965. [84] M. Thiele, J.Z.E. Soh, A. Knauer, D. Malsch, O. Stranik, R. Müller, A. Csáki, T.
[58] S. Kawasaki, K. Sue, R. Ookawar, Y. Wakashima, A. Suzuki, Y. Hakuta, K. Arai, Henkel, J.M. Köhler, W. Fritzsche, Gold nanocubes – direct comparison of
Engineering study of continuous supercritical hydrothermal method using a synthesis approaches reveals the need for amicrofluidic synthesis setup for
T-shaped mixer: experimental synthesis of NiO nanoparticles and CFD a high reproducibility, Chem. Eng. J. 288 (2016) 432–440.
simulation, J. Supercrit. Fluids 54 (2010) 96–102. [85] H.S. Santana, D.S. Tortola, J.L. Silva Jr., O.P. Taranto, Biodiesel synthesis in
[59] C.K. Chung, T.R. Shih, C.K. Chang, C.W. Lai, B.H. Wu, Design and experiments micromixer with static elements, Energy Convers. Manage. 141 (2017)
of a short-mixing-length baffled microreactor and its application to 28–39.
microfluidic synthesis of nanoparticles, Chem. Eng. J. 168 (2011) 790–798. [86] M. Thiele, A. Knauer, D. Malsch, A. Csáki, T. Henkel, J.M. Köhler, W. Fritzsche,
[60] J. Baumgard, A.M. Vogt, U. Kragl, K. Jahnisch, N. Steinfeldt, Application of Combination of microfluidic high-throughput production and parameter
microstructured devices for continuous synthesis of tailored platinum screening for efficient shaping of gold nanocubes using Dean-flow mixing,
nanoparticles, Chem. Eng. J. 227 (2013) 137–144. Lab Chip 17 (2017) 1487–1495.
[61] S. Bensaid, F.A. Deorsola, D.L. Marchisio, N. Russo, D. Fino, Flow field [87] A.C.S.N. Pessoa, C.C. Sipoli, L.G. de la Torre, Effects of diffusion and mixing
simulation and mixing efficiency assessment of the multi-inlet vortex mixer pattern on microfluidic-assisted synthesis of Chitosan/ATP nanoparticles,
for molybdenum sulfide nanoparticle precipitation, Chem. Eng. J. 238 (2014) Lab Chip 17 (2017) 2281–2293.
66–77. [88] R. Thiermann, R. Bleul, M. Maskos, Kinetic control of block copolymer
[62] S. Bensaid, M. Piumetti, C. Novara, F. Giorgis, A. Chiodoni, N. Russo, D. Fino, self-assembly in a micromixing device –mechanistical insight into vesicle
Catalytic oxidation of CO and soot over Ce-Zr-Pr mixed oxides synthesized formation process, Macromol. Chem. Phys. 218 (2017) 1600347.
in a multi-inlet vortex reactor: effect of structural defects on the catalytic [89] M. Bertrand, N. Lamarque, O. Lebaigue, E. Plasari, F. Ducros, Micromixing
activity, Nanoscale Res. Lett. 11 (2016) 494. characterisation in rapid mixing devices by chemical methods and LES
[63] H.S. Santana, J.L.S. Júnior, O.P. Taranto, Numerical simulation of mixing and modeling, Chem. Eng. J. 283 (2017) 462–475.
reaction of Jatropha curcas oil and ethanol for synthesis of biodiesel in [90] T. Lemenand, D. Della Valle, C. Habchi, H. Peerhossaini, Micromixing
micromixers, Chem. Eng. Sci. 132 (2015) 159–168. measurement by chemical probe in homogeneous and isotropic turbulence,
[64] Z.H. Tian, Y.J. Wang, J.H. Xu, G.S. Luo, Intensification of nucleation stage for Chem. Eng. J. 314 (2017) 453–465.
synthesizing high quality CdSe quantum dots by using preheated precursors [91] H. Qin, C. Zhang, Q. Xu, X. Dang, W. Li, K. Lei, L. Zhou, J. Zhang, Geometrical
in microfluidic devices, Chem. Eng. J. 302 (2016) 498–502. improvement of inline high shear mixers to intensify micromixing
[65] Z.H. Tian, J.H. Xu, Y.J. Wang, G.S. Luo, Microfluidic synthesis of performance, Chem. Eng. J. 319 (2017) 307–320.
monodispersed CdSe quantum dots nanocrystals by using mixed fatty [92] Y. Ouyang, Y. Xiang, H. Zou, G. Chu, J. Chen, Flow characteristics and
amines as ligands, Chem. Eng. J. 285 (2016) 20–26. micromixing modeling in a microporous tube-in-tube microchannel reactor
[66] Y. Qin, W. He, M. Su, Z. Fang, J. Gu, P. Ouyang, K. Guo, Continuous synthesis by CFD, Chem. Eng. J. 321 (2017) 533–545.
of ginkgolide B derivatives in a micro-flow system, Tetrahedron Lett. 57 [93] K. Oualha, M.B. Amar, A. Michau, A. Kanaev, Observation of cavitation in
(2016) 1243–1246. exocentric T-mixer, Chem. Eng. J. 321 (2017) 146–150.
[67] A. Polyzoidis, T. Altenburg, M. Schwarzer, S. Loebbecke, S. Kaskel, [94] H.C. Schwarzer, W. Peukert, Combined experimental/numerical study on the
Continuous microreactor synthesis of ZIF-8 with high space–time-yield and precipitation of nanoparticles, AIChE J. 50 (2004) 3234–3247.
tunable particle size, Chem. Eng. J. 283 (2016) 971–977. [95] H.C. Schwarzer, F. Schwertfirm, M. Manhart, H.J. Schmid, d.W. Peukert,
[68] S. Zhu, Y. Lu, R. Faust, Micromixing enhanced synthesis of HRPIBs catalyzed Predictive simulation of nanoparticle precipitation based on the population
by EADC/bis(2-chloroethyl)ether complex, RSC Adv. 7 (2017) 27629–27636. balance equation, Chem. Eng. Sci. 61 (2006) 167–181.
[69] N. Daito, N. Aoki, J. Yoshida, K. Mae, Selective condensation reaction of [96] J. Gradl, H. Schwarzer, F. Schwertfirmc, M. Manhart, W. Peukert,
phenols and hydroxybenzyl alcohol using micromixers based on collision of Precipitation of nanoparticles in a T-mixer: coupling the particle population
fluid segments, Ind. Eng. Chem. Res. 45 (2006) 4954–4961. dynamics with hydrodynamics through direct numerical simulation, Chem.
[70] M. Luty-Błocho, K. Fitzne, V. Hesse, P. Lobb, M. Maskos, D. Metzke, K. Eng. Process. 45 (2006) 908–916.
Pacławski, M. Wojnicki, Synthesis of gold nanoparticles in an interdigital [97] H. Kawanami, K. Matsushim, M. Sato, Y. Ikushim, Rapid and highly selective
micromixer using ascorbic acid and sodium borohydride as reducers, Chem. copper-free sonogashira coupling in high-pressure, high-temperature water
Eng. J. 171 (2011) 279–290. in a microfluidic system, Angew. Chem. Int. Ed. 46 (2007) 5129–5132.
[71] Y.J. Jung, S.H. Park, K.H. Song, J. Choe, Recrystallization of polyethylene [98] Y.T. Chen, K.H. Chen, W.F. Fang, S.H. Tsai, J.M. Fang, J.T. Yang, Flash synthesis
submicron particles using a continuous flow micromixer system, Powder of carbohydrate derivatives in chaotic microreactors, Chem. Eng. J. 174
Technol. 217 (2012) 325–329. (2011) 421–424.
[72] D.V.R. Kumar, B.L.V. Prasad, A.A. Kulkarni, Segmented flow synthesis of Ag [99] W.F. Fang, J.T. Yang, A novel microreactor with 3D rotating flow to boost
nanoparticles in spiral microreactor: role of continuous and disperzsed fluid reaction and mixing of viscous fluids, Sens. Actuators B: Chem. 140
phase, Chem. Eng. J. 192 (2012) 357–368. (2009) 629–642.
[73] D. Yamamoto, T. Maki, S. Watanabe, H. Tanaka, M.T. Miyahara, K. Mae, [100] D. Lee, Y.T. Chen, Mixing in tangentially crossing microchannels, AIChE J. 57
Synthesis and adsorption properties of ZIF-8 nanoparticles using a (2011) 571–580.
micromixer, Chem. Eng. J. 227 (2013) 145–150. [101] S. Mozharov, A. Nordon, D. Littlejohn, C. Wiles, P. Watts, P. Dallin, J.M.
[74] S. Haroun, L. Wang, T.J. Ruth, P.C.H. Li, Computational fluid dynamics study Girkin, Improved method for kinetic studies in microreactors using flow
of the synthesis process for a PET radiotracer compound, [11 C]raclopride on manipulation and noninvasive Raman spectrometry, J. Am. Chem. Soc. 133
a microfluidic chip, Chem. Eng. Process. 70 (2013) 140–147. (2011) 3601–3608.
[75] T. Ono, K. Sue, D. Furuta, M. Aoki, Y.a Hakuta, Y. Takebayashi, S. Yoda, T. [102] Y. Han, J.J. Wang, X.P. Gu, L.F. Feng, Numerical simulation on micromixing of
Furuya, T. Sato, T. Hiak, Continuous hydrothermal synthesis of Ca1−x Srx TiO3 viscous fluids in a stirred-tank reactor, Chem. Eng. Sci. 74 (2012) 9–17.
solid-solution nanoparticles using a T-type micromixer, J. Supercrit. Fluids [103] X.X. Duan, X. Feng, C. Yang, Z.-S. Mao, Numerical simulation of micro-mixing
85 (2014) 159–164. in stirred reactors using the engulfment model coupled with CFD, Chem.
[76] A.S. Yang, F.C. Chuang, C.K. Chen, M.H. Lee, S.W. Chen, T.-L. Su, Y.C. Yang, A Eng. Sci. 140 (2016) 179–188.
high-performance micromixer using three-dimensional Tesla structures for [104] M.J. Hossain, H. Tsunoyama, M. Yamauchi, N. Ichikuni, T. Tsukuda,
bio-applications, Chem. Eng. J. 263 (2015) 444–451. High-yield synthesis of PVP-stabilized small Pt clusters by microfluidic
[77] Q. Deng, R. Shen, Z. Zhao, M. Yan, L. Zhang, The continuous flow synthesis of method, Catal. Today 183 (2012) 101–107.
2,4,5-trifluorobenzoic acid via sequential Grignard exchange and [105] N.T. Zuidhof, M.H.J.M.D. Croon, J.C. Schouten, J.T. Tinge, Beckmann
carboxylation reactions using microreactors, Chem Eng. J. 262 (2015) rearrangement of cyclohexanone oxime to ε-caprolactam in a microreactor,
1168–1174. Chem. Eng. Technol. 35 (2012) 1257–1261.
[78] M. Maeki, T. Saito, Y. Sato, T. Yasui, N. Kaji, A. Ishida, H. Tani, Y. Baba, H. [106] N.T. Zuidhof, M.H.J.M.D. Croon, J.C. Schouten, J.T. Tinge, Beckmann
Harashima, M. Tokeshi, A strategy for synthesis of lipid nanoparticles using rearrangement of cyclohexanone oxime in a microreactor setup with
microfluidic devices with a mixer structure, RSC Adv. 5 (2015) internal recirculation, Chem. Eng. Technol. 26 (2013) 1387–1394.
46181–46185. [107] P. Plouffe, D.M. Roberge, A. Macchi, Liquid–liquid flow regimes and mass
[79] H.S. Santana, J.L.S. Júnior, O.P. Taranto, Numerical simulations of biodiesel transfer in various micro-reactors, Chem. Eng. J. 300 (2016) 9–19.
synthesis in microchannels with circular obstructions, Chem. Eng. Process. [108] E. Mielke, D.M. Roberge, A. Macchi, Microreactor mixing-unit design for fast
98 (2015) 137–146. liquid–liquid reactions, J. Flow Chem. 6 (2016) 279–287.
700 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

[109] P. Plouffe, M. Bittel, J. Sieber, D.M. Roberge, A. Macchi, On the scale-up of [136] C. Tonhauser, D. Wilms, F. Wurm, E. Berger-Nicoletti, M. Maskos, H. Löwe, H.
micro-reactors for liquid–liquid reactions, Chem. Eng. Sci. 143 (2016) Frey, Multihydroxyl-functional polystyrenes in continuous flow,
216–225. Macromolecules 43 (2010) 5582–5588.
[110] J. Choe, J.H. Seo, Y. Kwon, K.H. Song, Lithium–halogen exchange reaction [137] J. Morsbach, A.H.E. Müller, E. Berger-Nicoletti, H. Frey, Living polymer chains
using microreaction technology, Chem. Eng. J. 135S (2008) S17–S20. with predictable molecular weight and dispersity via carbanionic
[111] Y. Ukita, T. Asano, K. Fujiwara, K. Matsui, M. Takeo, S. Negoro, T. Kanie, M. polymerization in continuous flow: mixing rate as a key parameter,
Katayama, Y. Utsumi, Application of vertical microreactor stack with Macromolecules 49 (2016) 5043–5050.
polystylene microbeads to immunoassay, Sens. Actuators A: Phys. 145–146 [138] L. Qiu, K. Wang, S. Zhu, Y. Lu, G. Luo, Kinetics study of acrylic acid
(2008) 449–455. polymerization with a microreactor platform, Chem. Eng. J. 284 (2016)
[112] W.F. Fang, J.T. Yang, A novel microreactor with 3D rotating flow to boost 233–239.
fluid reaction and mixing of viscous fluids, Sens. Actuators B: Chem. 140 [139] N. Zhu, Z. Zhang, W. Feng, Y. Zeng, Z. Li, Z. Fang, K. Zhang, Z. Li, K. Guo,
(2009) 629–642. Sn(OTf)2 catalyzed continuous flow ring-opening polymerization of
[113] D. Ziegenbalg, P. Löb, M. Al-Rawashdeh, D. Kralisch, V. Hessel, F. Schönfeld, ␧-caprolactone, RSC Adv. 5 (2015) 31554–31557.
Use of ‘smart interfaces’ to improve the liquid-sided mass transport in a [140] A. Nagaki, K. Kawamura, S. Suga, T. Ando, M. Sawamoto, J. Yoshida, Cation
falling film microreactor, Chem. Eng. Sci. 65 (2010) 3557–3566. pool-initiated controlled/living polymerization using microsystems, J. Am.
[114] M.H. Hsu, W.F. Fang, Y.H. Lai, J.T. Yang, T.L. Tsai, D.B. Shieh, Enhanced mobile Chem. Soc. 126 (2004) 14702–14703.
hybridization of gold nanoparticles decorated with oligonucleotide in [141] T. Iwasaki, J. Yoshida, Free radical polymerization in microreactors
microchannel devices, Lab Chip 10 (2010) 2583–2587. Significant improvement in molecular weight distribution control,
[115] Y. Ko, J. Maeng, Y.n Ahn, S. Hwang, DNA ligation using a disposable Macromolecules 38 (2005) 1159–1163.
microfluidic device combined with a micromixer and microchannel reactor, [142] U. Yasuhiro, M.A.Y. Yoichi, B. Tomohiko, F. Naoshi, U. Masaharu, K. Takehiko,
Sens. Actuators B: Chem. 157 (2011) 735–741. Instantaneous carbon–carbon bond formation using a microchannel reactor
[116] T. Yasukawa, W. Ninomiya, K. Ooyachi, N. Aoki, K. Mae, Enhanced with a catalytic membrane, J. Am. Chem. Soc. 128 (2006) 15994–15995.
production of ethyl pyruvate using gas–liquid slug flow in microchannel, [143] T. Iwasaki, J. Yoshida, CF3 SO3 H initiated cationic polymerization of
Chem. Eng. J. 167 (2011) 527–530. diisopropenylbenzenes in macrobatch and microflow systems, Macromol.
[117] Y.T. Chen, W.F. Fang, Y.C. Liu, J.T. Yang, Analysis of chaos and FRET reaction Rapid Commun. 28 (2007) 1219–1224.
in split-and-recombine microreactors, Microfluid. Nanofluid. 11 (2011) [144] C. Rosenfeld, C. Serra, C. Brochon, G. Hadziioannou, Influence of micromixer
339–352. characteristics on polydispersity index of block copolymers synthesized in
[118] Y. Li, D. Zhang, X. Feng, Y. Xu, B.g Liu, A microsecond microfluidic mixer for continuous flow microreactors, Lab Chip 8 (2008) 1682–1687.
characterizing fast biochemical reactions, Talanta 88 (2012) 175–180. [145] A. Nagaki, Y. Tomida, J. Yoshida, Microflow-system-controlled anionic
[119] M. Al-Rawashdeh, A. Cantu-Perec, D. Ziegenbalg, P. Löb, A. Gavriilidis, V. polymerization of styrenes, Macromolecules 41 (2008) 6322–6330.
Hessel, F. Schonfeld, Microstructure-based intensification of a falling film [146] A. Nagaki, Y. Tomida, A. Miyazaki, J. Yoshida, Microflow system controlled
microreactor through optimal film setting with realistic profiles and anionic polymerization of alkyl methacrylates, Macromolecules 42 (2009)
in-channel induced mixing, Chem. Eng. J. 179 (2012) 318–329. 4384–4387.
[120] X. Guo, Y. Fan, L. Luo, Mixing performance assessment of a multi-channel [147] K. Iida, T.Q. Chastek, K.L. Beers, K.A. Cavicchi, J. Chun, M.J. Fasolka, Living
mini heat exchanger reactor with arborescent distributor and collector, anionic polymerization using a microfluidic reactor, Lab Chip 9 (2009)
Chem. Eng. J. 227 (2013) 116–127. 339–345.
[121] M. Sudar, Z. Findrik, D. Vasć-Raĉki, P. Clapés, C. Lozano, Aldol addition of [148] A. Nagaki, A. Miyazaki, J. Yoshida, Synthesis of polystyrenes-poly(alkyl
dihydroxyacetone to N-Cbz-3-aminopropanal catalyzed by two aldolases methacrylates) block copolymers via anionic polymerization using an
variants in microreactors, Enzyme Microb. Technol. 53 (2013) 38–45. integrated flow microreactor system, Macromolecules 43 (2010)
[122] C. Dong, J.S. Zhang, K. Wang, G.S. Luo, Micromixing performance of 8424–8429.
nanoparticle suspensions in a micro-sieve dispersion reactor, Chem. Eng. J. [149] A. Nagaki, A. Miyazaki, Y. Tomida, J. Yoshida, Anionic polymerization of alkyl
253 (2014) 8–15. methacrylates using flow microreactor systems, Chem. Eng. J. 167 (2011)
[123] Z. Liu, L. Guo, T. Huang, L. Wen, J. Chen, Experimental and CFD studies on the 548–555.
intensified micromixing performance of micro-impinging stream reactors [150] A. Nagaki, Y. Nakahara, K. Akahori, J. Yoshida, Living anionic polymerization
built from commercial T-junctions, Chem. Eng. Sci. 119 (2014) 124–133. of tert-butyl acrylate in a flow microreactor system and its applications to
[124] S. Mitic, J.W. van Nieuwkasteele, A. van den Berg, S. de Vries, Design of the synthesis of block copolymers, Macromol. React. Eng. 6 (2012) 467–472.
turbulent tangential micro-mixers that mix liquids on the nanosecond time [151] J. Vandenbergh, T. de Moraes Ogawa, T. Junkers, Precision synthesis of
scale, Anal. Biochem. 469 (2015) 19–26. acrylate multiblock copolymers from consecutive microreactor RAFT
[125] C. Dong, K. Wang, J.S. Zhang, G.S. Luo, Reaction kinetics of cyclohexanone polymerizations, J. Polym. Sci. A 51 (2013) 2366–2374.
ammoximation over TS-1 catalyst in a microreactor, Chem. Eng. Sci. 126 [152] L.S. Méndez-Portillo, C. Dubois, P.A. Tanguy, Free-radical polymerization of
(2015) 633–640. styrene using a split-and-recombination(SAR) and multilamination
[126] P. Plouffe, D.M. Roberge, J. Sieber, M. Bittel, A. Macchi, Liquid–liquid mass microreactors, Chem. Eng. J. 256 (2014) 212–221.
transfer in a serpentine micro-reactor using various solvents, Chem. Eng. J. [153] J.J. Haven, J. Vandenbergh, T. Junkers, Watching polymers grow: real time
285 (2016) 605–615. monitoring of polymerizations via an on-line ESI–MS/microreactor
[127] S. Watanabe, S. Ohsaki, T. Hanafusa, K. Takada, H. Tanaka, K. Mae, M.T. coupling, Chem. Commun. 51 (2015) 4611–4614.
Miyahara, Synthesis of zeolitic imidazolate framework-8 particles of [154] Y. Zhang, N.E. Benes, R.G.H. Lammertink, Visualization and characterization
controlled sizes shapes, and gate adsorption characteristics using a central of interfacial polymerization layer formation, Lab. Chip 15 (2015) 575–580.
collision-type microreactor, Chem. Eng. J. 313 (2017) 724–733. [155] A. Nagaki, M. Takumi, Y. Tani, J. Yoshida, Polymerization of vinyl ethers
[128] I.P.R. Grundtvig, A.E. Daugaard, J.M. Woodley, K.V. Gernaey, U. Krühne, initiated by dendritic cations using flow microreactors, Tetrahedron 71
Shape optimization as a tool to design biocatalytic microreactors, Chem. (2015) 5973–5978.
Eng. J. 322 (2017) 215–223. [156] Y. Tani, M. Takumi, S. Moronaga, A. Nagaki, J. Yoshida, Flash cationic
[129] M. Rahimi, F. Mohammadi, M. Basiri, M.A. Parsamoghadam, M.M. Masahi, polymerization followed by bis-endfunctionalization. A new approach to
Transesterification of soybean oil in four-way micromixers for biodiesel linear-dendritic hybrid polymers, Eur. Polym. J. 80 (2016) 227–233.
production using a cosolvent, J. Taiwan Inst. Chem. Eng. 64 (2016) 203–210. [157] A. Nagaki, Y. Nakahara, M. Furusawa, T. Sawaki, T. Yamamoto, H. Toukairin,
[130] J. Wang, F. Zhang, Y. Wang, G. Luo, W. Cai, A size-controllable preparation S. Tadokoro, T. Shimazaki, T. Ito, M. Otake, H. Arai, N. Toda, K. Ohtsuka, Y.
method for indium tin oxide particles using a membrane dispersion Takahashi, Y. Moriwaki, Y. Tsuchihashi, K. Hirose, J. Yoshida, Feasibility
micromixer, Chem. Eng. J. 293 (2016) 1–8. study on continuous flow controlled/living anionic polymerization
[131] W. Liu, Z. Xu, L. Sun, P. Guo, C. Zeng, C. Wang, L. Zhang, processes, Org. Process Res. Dev. 20 (2016) 1377–1382.
Polymerization-induced phase separation fabrication: a versatile [158] M. Ryu, J.A. Kimber, T. Sato, R. Nakatani, T. Hayakawa, M. Romano, C.
microfluidic technique to prepare microfibers with various cross sectional Pradere, A.A. Hovhannisyan, S.G. Kazarian, J. Morikawa, Infrared
shapes and structures, Chem. Eng. J. 315 (2017) 25–34. thermo-spectroscopic imaging of styrene radical polymerization in
[132] T. Honda, M. Miyazaki, H. Nakamura, H. Maeda, Controllable polymerization microfluidics, Chem. Eng. J. 324 (2017) 259–265.
of N-carboxy anhydrides in a microreaction system, Lab Chip 5 (2005) [159] K. Miyabayashi, O. Tonomura, S. Hasebe, Estimation of gas and liquid slug
812–818. lengths for T-shaped microreactors, Chem. Eng. J. 262 (2015) 1137–1143.
[133] C. Rosenfeld, C. Serra, C. Brochon, V. Hessel, G.S. Hadziioannou, Use of [160] I.V. Gürsel, S.K. Kurt, J. Aalders, Q. Wang, T. Noël, K.D.P. Nigam, N. Kockmann,
micromixers to control the molecular weight distribution in continuous V. Hessel, Utilization of milli-scale coiled flow inverter in combination with
two-stage nitroxide-mediated copolymerizations, Chem. Eng. J. 135S (2008) phase separator for continuous flow liquid–liquid extraction processes,
S242–S246. Chem. Eng. J. 283 (2016) 855–868.
[134] D. Kessler, H. Löwe, P. Theato, Synthesis of defined poly(silsesquioxane)s: [161] N. Azimi, M. Rahimi, Magnetic nanoparticles stimulation to enhance
fast polycondensation of trialkoxysilanes in a continuous-flow microreactor, liquid–liquid two-phase mass transfer under static and rotating magnetic
Macromol. Chem. Phys. 210 (2009) 807–813. fields, J. Magn. Magn. Mater. 422 (2017) 188–196.
[135] S. Kundu, A.S. Bhangale, W.E. Wallace, K.M. Flynn, C.M. Guttman, R.A. Gross, [162] S. Dziomba, M. Araya-Farias, C. Smadja, M. Taverna, B. Carbonnier, N.T. Tran,
K.L. Beers, Continuous flow enzyme catalyzed polymerization in a Solid supports for extraction and preconcentration of proteins and peptides
microreactor, J. Am. Chem. Soc. 133 (2011) 6006–6011. in microfluidic devices: a review, Anal. Chim. Acta 955 (2017) 1–26.
C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702 701

[163] K. Mae, T. Maki, I. Hasegawa, U. Eto, Y. Mizutani, N. Honda, Development of a micro-unit operations and online ICP-MS determination, Talanta 139 (2015)
new micromixer based on split/recombination for mass production and its 123–131.
application to soap free emulsifier, Chem. Eng. J. 101 (2004) 31–38. [192] M. Heule, A. Manz, Sequential DNA hybridisation assays by fast
[164] T. Sprogies, J.M. Köhler, G.A. Groß, Evaluation of static micromixers for micromixing, Lab Chip 4 (2004) 506–511.
flow-through extraction by emulsification, Chem. Eng. J. 135S (2008) [193] X. Chen, D. Cui, C. Liu, H. Li, J. Chen, Continuous flow microfluidic device for
S199–S202. cell separation, cell lysis and DNA purification, Anal. Chim. Acta 584 (2007)
[165] A. Ufer, D. Sudhoff, A. Mescher, D.W. Agar, Suspension catalysis in a 237–243.
liquid–liquid capillary microreactor, Chem. Eng. J. 167 (2011) 468–474. [194] Y. Huh, J. Choi, T. Park, Y. Hong, W. Hong, S. Lee, Microfluidic cell disruption
[166] T. Vandermeersch, R. Goovaerts, J. Luyten, J.F.M. Denayer, W.D. Malsche, system employing a magnetically actuated diaphragm, Electrophoresis 25
Tracking the liquid–liquid extraction performance in mesoflow reactors, (2007) 4748–4757.
Chem. Eng J. 279 (2015) 9–17. [195] Y. Okabe, Y. Chen, R. Purohit, R.M. Corn, A.P. Lee, Piezoelectrically driven
[167] T. Vandermeersch, L. Gevers, W.D. Malsche, A robust multistage mesoflow vertical cavity acoustic transducers for the convective transport and rapid
reactor for liquid–liquid extraction for the separation of Co/Ni with cyanex detection of DNA and protein binding to DNA microarrays with SPR imaging
272, Sep. Purif. Technol. 168 (2016) 32–38. – a parametric study, Biosens. Bioelectron. 35 (2012) 37–43.
[168] C.L. Lin, W.H. Chang, C.H. Wang, C.H. Lee, T.Y. Chen, F.J. Jan, G.B. Lee, A [196] F. Nason, E. Morganti, C. Collini, C. Ress, S. Bersini, G. Pennati, F. Boschetti, A.
microfluidic system integrated with buried optical fibers for detection of Colombini, G. Lombardi, G. Banfi, L. Lorenzelli, G. Dubini, Design of
Phalaenopsis orchid pathogens, Biosens. Bioelectron. 63 (2016) 572–579. microfluidic devices for drug screening on in-vitro cells for osteoporosis
[169] C.H. Wang, C.P. Chang, G.B. Lee, Integrated microfluidic device using a single therapies, Microelectron. Eng. 88 (2011) 1801–1806.
universal aptamer to detect multiple types of influenza viruses, Biosens. [197] L. Wang, R. Kodzius, X. Yi, S. Li, Y. Hui, W. Wen, Prototyping chips in
Bioelectron. 86 (2016) 247–254. minutes: direct Laser Plotting (DLP) of functional microfluidic structures,
[170] C. Yao, K. Zhu, Y. Liu, H. Liu, F. Jiao, G. Chen, Intensified CO2 absorption in a Sens. Actuators B: Chem. 168 (2012) 214–222.
microchannel reactor under elevated pressures, Chem. Eng. J. 319 (2017) [198] T.A. Balbino, A.R. Azzoni, L.G. de la Torre, Microfluidic devices for continuous
179–190. production of pDNA/cationic liposome complexes for gene delivery and
[171] C.H. Lin, Y.N. Wang, L.M. Fu, Integrated microfluidic chip for rapid DNA vaccine therapy, Colloids Surf. B: Biointerfaces 111 (2013) 203–210.
digestion and time-resolved capillary electrophoresis analysis, [199] Y.H. Lin, C.C. Wang, K.F. Lei, Bubble-driven mixer integrated with a
Biomicrofluidics 6 (2012) 012818. microfluidic bead-based ELISA for rapid bladder cancer biomarker detection,
[172] L.M. Fu, C.H. Lin, A rapid DNA digestion system, Biomed. Microdevices 9 Biomed. Microdevices 16 (2014) 199–207.
(2007) 277–286. [200] G.R. Aguirre, V. Efremov, M. Kitsara, J. Ducrée, Integrated micromixer for
[173] I.N. Kefala, V.E. Papadopoulos, G. Karpou, G. Kokkoris, G. Papadakis, A. incubation and separation of cancer cells on a centrifugal platform using
Tserepi, A labyrinth split and merge micromixer for bioanalytical inertial and dean forces, Microfluid. Nanofluid. 18 (2015) 513–526.
applications, Microfluid. Nanofluid. 19 (2015) 1047–1059. [201] P. Novoa, M.D. Aicaa, D. Janaseka, R.P. Zahedia, High temporal resolution
[174] A.S. Kastania, K. Tsougeni, G. Papadakis, E. Gizeli, G. Kokkoris, A. Tserepi, E. study of phosphorylation events in HEK cells using a micromixer
Gogolides, Plasma micro-nanotextured polymeric micromixer for DNA microfluidic device, Procedia Eng. 168 (2016) 1430–1433.
purification with high efficiency and dynamic range, Anal. Chim. Acta 942 [202] Q. Lang, Y. Ren, D. Hobson, Y. Tao, L. Hou, Y. Jia, Q. Hu, J. Liu, X. Zhao, In-plane
(2016) 58–67. microvortices micromixer-based AC electrothermal for testing drug induced
[175] V.E. Papadopoulos, I.N. Kefalaa, G. Kaprou, G. Kokkoris, D. Moschou, G. death of tumor cells, Biomicrofluidics 10 (2016) 064102.
Papadakis, E. Gizeli, A. Tserepi, A passive micromixer for enzymatic [203] T.Y. Lee, K.A. Hyun, S.I. Kim, H.I. Jung, An integrated microfluidic chip for
digestion of DNA, Microelectron. Eng. 124 (2014) 42–46. one-step isolation of circulating tumor cells, Sens. Actuators B: Chem. 238
[176] N.Y. Lee, M. Yamada, M. Seki, Development of a passive micromixer based (2017) 1144–1150.
on repeated fluid twisting and flattening, and its application to DNA [204] M. Hashimoto, F. Barany, F. Xu, S.A. Soper, Serial processing of biological
purification, Anal. Bioanal. Chem. 5 (2005) 776–782. reactions using flow-through microfluidic devices: coupled PCR/LDR for the
[177] H.Y. Lee, J. Voldman, Optimizing micromixer design for enhancing detection of low-abundant DNA point mutations, Analyst 132 (2007)
dielectrophoretic microconcentrator performance, Anal. Chem. 79 (2007) 913–921.
1833–1839. [205] S.J. Kim, F. Wang, M.A. Burns, K. Kurabayashi, Temperature-programmed
[178] J.N. Kuo, Y.H. Zhan, Microfluidic chip for rapid and automatic extraction of natural convection for micromixing and biochemical reaction in a single
plasma from whole human blood, Microsyst. Technol. 21 (2015) 255–261. microfluidic chamber, Anal. Chem. 81 (2009) 4510–4516.
[179] N. Azimi, M. Rahimi, N. Abdollahi, Using magnetically excited nanoparticles [206] J. Choi, Y. Kim, J. Byun, J. Ahn, S. Chung, D. Gweon, M. Kimd, T.S. Seo, An
for liquid–liquid two-phase mass transfer enhancement in a Y-type integrated allele-specific polymerase chain reaction-microarray chip for
micromixer, Chem. Eng. Process. 97 (2015) 12–22. multiplex single nucleotide polymorphism typing, Lab Chip 12 (2012)
[180] C.C. Domínguez, T. Gamse, Process intensification by the use of micro 5146–5154.
devices for liquid fractionation with supercritical carbondioxide, Chem. Eng. [207] C.M. Chang, L.F. Chiu, Y.H. We, D.B. Shieh, G.B. Lee, Integrated
Res. Des. 108 (2016) 139–145. three-dimensional system-on-chip for direct quantitative detection of
[181] T.T. Shih, C.C. Hsieh, Y.T. Luo, Y.A. Su, P.H. Chen, Y.C. Chuang, Y.C. Sun, A mitochondrial DNA mutationin affected cells, Biosens. Bioelectron. 48
high-throughput solid-phase extraction microchip combined with (2013) 6–11.
inductively coupled plasma-mass spectrometry for rapid determination of [208] M. Geissler, K. Li, X. Zhang, L. Clime, G.P. Robideau, G.J. Bilodeau, T. Veres,
trace heavy metals in natural water, Anal. Chim. Acta 916 (2016) 24–32. Integrated air stream micromixer for performing bioanalytical assays on a
[182] O. Jafari, M. Rahimi, F.H. Kakavandi, Liquid–liquid extraction in twisted plastic chip, Lab Chip 14 (2014) 3750–3761.
micromixers, Chem. Eng. Process. 101 (2016) 33–40. [209] A.A. Sayad, F. Ibrahim, S.M. Uddin, K.X. Pei, M.S. Mohktar, M. Madou, K.L.
[183] F.H. Kakavandi, M. Rahimi, O. Jafari, N. Azimi, Liquid–liquid two-phase mass Thong, A microfluidic lab-on-a-disc integrated loop mediated isothermal
transfer in T-type micromixers with different junctions and cylindrical pits, amplification for foodborne pathogen detection, Sens. Actuators B: Chem.
Chem. Eng. Process. 107 (2016) 58–67. 227 (2016) 600–609.
[184] M. Rahimi, O. Jafari, A. Mohammdifar, Intensification of liquid–liquid mass [210] S.L. Chen, W.H. Chang, C.H. Wang, H.L. You, J.J. Wu, T.H. Liu, M.S. Lee, G.B.
transfer in micromixer assisted by ultrasound irradiation and Fe3 O4 Lee, An integrated microfluidic system for live bacteria detection from
nanoparticles, Chem. Eng. Process 111 (2017) 79–88. human joint fluid samples by using ethidium monoazide and loop-mediated
[185] G. Peng, Q. He, Y. Lu, J. Huang, J. Lin, Flow injection microfluidic device with isothermal amplification, Microfluid. Nanofluid. 21 (2017) 87.
on-line fluorescent derivatization for the determination of Cr(III) and Cr(VI) [211] J.A. Lounsbury, A. Karlsson, D.C. Miranian, S.M. Cronk, D.A. Nelson, J. Li, D.M.
in water samples after solid phase extraction, Anal. Chim. Acta 955 (2017) Haverstick, P. Kinnon, D.J. Saul, J.P. Landers, From sample to PCR product in
58–66. under 45 minutes: a polymeric integrated microdevice for clinical
[186] C.C. Domínguez, T. Gamse, Utilization of micro-mixers for supercritical fluid andforensic DNA analysis, Lab Chip 13 (2013) 1384–1393.
fractionation: Influence of the residence time, J. Supercrit. Fluids 132 (2018) [212] M.C. Giuffrida, G. Spoto, Integration of isothermal amplification methods in
17–23. microfluidic devices: recent advances, Biosens. Bioelectron. 90 (2017)
[187] H.M. Ji, V. Samper, Y. Chen, W.C. Hui, H.J. Lye, F.B. Mustafa, A.C. Lee, L. Cong, 174–186.
C.K. Heng, T.M. Lim, DNA purification silicon chip, Sens. Actuators A: Phys. [213] Y. Li, F. Xu, C. Liu, Y. Xu, X. Feng, B.F. Liu, A novel microfluidic mixer based on
139 (2007) 139–144. dual-hydrodynamic focusing for interrogating the kinetics of DNA–protein
[188] M.C. Morales, J.D. Zahn, Droplet enhanced microfluidic-based DNA interaction, Analyst 138 (2013) 4475–4482.
purification from bacterial lysates via phenol extraction, Microfluid. [214] W.B. Lee, Y.H. Chen, H.I. Lin, S.C. Shiesh, G.B. Lee, An integrated microfluidic
Nanofluid. 9 (2010) 1041–1049. system for fast, automatic detection of C-reactive, Protein Sens. Actuators B:
[189] C.H. Tai, J.W. Shin, T.Y. Chang, S.K. Hsiung, C.C. Lin, G.B. Lee, An integrated Chem. 157 (2011) 710–721.
microfluidic system capable of sample pretreatment and hybridization for [215] Y.N. Yang, H.I. Lin, J.H. Wang, S.C. Shiesh, G.B. Lee, An integrated microfluidic
microarrays, Microfluid. Nanofluid. 10 (2011) 999–1009. system for C-reactive protein measurement, Biosens. Bioelectron. 24 (2009)
[190] P.C. Chen, An evaluation of a real-time passive micromixer to the 3091–3096.
performance of a continuous flow type microfluidic reactor, BioChip J. 7 [216] M.X. Lin, K.A. Hyun, H. Moon, T. Sim, J. Lee, J. Park, S. Lee, H. Jung, Continuous
(2013) 227–233. labeling of circulating tumor cells with microbeads using a vortex
[191] G. Hellé, C. Mariet, G. Cote, Liquid–liquid extraction of uranium(VI) with micromixer for highly selective isolation, Biosens. Bioelectron. 40 (2013)
Aliquaté 336 from HCl media in microfluidic devices: combination of 63–67.
702 C.-Y. Lee, L.-M. Fu / Sensors and Actuators B 259 (2018) 677–702

[217] N. Rajabi, J. Bahnemann, T. Tzeng, O. Barradas, A. Zeng, J. MülleraInstitute, [243] K.Y. Lee, S. Park, Y.R. Lee, S.K. Chung, Magnetic droplet microfluidic system
Lab-on-a-chip for cell perturbation lysis, and efficient separation of incorporated with acoustic excitation for mixing enhancement, Sens.
sub-cellular components in a continuous flow mode, Sens. Actuators A: Actuators A: Phys. 243 (2016) 59–65.
Phys. 215 (2014) 136–143. [244] X. Chen, Ca.L. Ren, A microfluidic chip integrated with droplet generation,
[218] T. Femmer, M.L. Eggersdorfer, A.J.C. Kuehne, M. Wessling, Efficient pairing, trapping, merging, mixing and releasing, RSC Adv. 7 (2017)
gas–liquid contact using microfluidic membrane devices with staggered 16738–16750.
herringbone mixers, Lab Chip 15 (2015) 3132–3137. [245] A. Larrea, A. Clemente, E. Luque-Michel, V. Sebastian, Efficient production of
[219] J. Oh, S.C. Choi, D. Lee, K.I. Kang, J. Jung, Interactive control of micro-mixing hybrid bio-nanomaterials by continuous microchannel emulsification:
and in-situ gelation of hydrogels for cell-laden cylindrical microgel dye-doped SiO2 and Au-PLGA nanoparticles, Chem. Eng. J. 316 (2017)
fabrication, Dig. J. Nanomater. Biostruct. 11 (2016) 329–336. 663–672.
[220] A. Cosentino, H. Madadi, P. Vergara, R. Vecchione, F. Causa, P.A. Netti, An [246] E. Samiei, M.D. de Leon Derby, A. Van den Berg, M. Hoorfar, An
efficient planar accordion-shaped micromixer: from biochemical mixing to electrohydrodynamic technique for rapid mixing in stationary droplets on
biological application, Sci. Rep. 5 (2015) 17876. digital microfluidic platforms, Lab. Chip 17 (2017) 227–234.
[221] X. Jia, W. Wang, Q. Han, Z. Wang, Y. Jia, Z. Hu, Micromixer based preparation [247] H.S. Abadi, M. Packirisamy, R. Wüthrich, High performance cascaded PDMS
of functionalized liposomes and targeting drug delivery, ACS Med. Chem. micromixer based on split-and-recombination flows for lab-on-a-chip
Lett. 7 (2016) 429–434. applications, RSC Adv. 3 (2013) 7296–7305.
[222] M. Cho, S. Chung, Y.T. Kim, J.H. Jung, D.H. Kim, T.S. Seo, A fully integrated [248] M. Bezagu, S. Arseniyadis, J. Cossy, O. Couture, M. Tanter, F. Monti, P.
microdevice for biobarcode assay based biological agent detection, Lab Chip Tabeling, A fast and switchable microfluidic mixer based on
15 (2015) 2744–2748. ultrasound-induced vaporization of perfluorocarbon, Lab. Chip 15 (2015)
[223] Y. Gao, P. Tran, K. Petkovic-Duran, T. Swallow, Y. Zhu, Acoustic micromixing 2025–2029.
increases antibody-antigen binding in immunoassays, Biomed. [249] A. Bohr, J. Boetker, Y. Wang, H. Jensen, J. Rantanen, M. Beck-Broichsitter,
Microdevices 17 (2015) 79. High-throughput fabrication of nanocomplexes using 3D-printed
[224] S.Y. Yang, K.Y. Lien, K.J. Huang, H.Y. Lei, G.B. Lee, Micro flow cytometry micromixers, J. Pharm. Sci. 106 (2017) 835–842.
utilizing a magnetic bead-based immunoassay for rapid virus detection, [250] G. Santillo, F.A. Deorsola, S. Bensaid, N. Russo, D. Fino, MoS2 nanoparticle
Biosens. Bioelectron. 24 (2008) 855–862. precipitation in turbulent micromixers, Chem. Eng. J. 207–208 (2012)
[225] Y.F. Lee, K.Y. Lien, H.Y. Lei, G.B. Lee, An integrated microfluidic system for 322–328.
rapid diagnosis of dengue virus infection, Biosens. Bioelectron. 25 (2009) [251] F.T.G. van den Brink, T. Wigger, L. Ma, M. Odijk, W. Olthuis, U. Karstb, A. van
745–752. den Berg, Oxidation and adduct formation of xenobiotics in a microfluidic
[226] Z. Wang, T. Han, T.n Jeon, S. Park, S. Kim, Rapid detection and quantification electrochemical cell with boron doped diamond electrodes and an
of bacteria using an integrated micro/nanofluidic device, Sens. Actuators B: integrated passive gradient rotation mixer, Lab. Chip 16 (2016) 3990–4001.
Chem. 178 (2013) 683–688. [252] F.G. Ergin, B.B. Watz, K. Ērglis, A. Cēbers, Time-resolved velocity
[227] B. Zhang, Y. Luo, B. Zhou, Q. Wang, P.D. Millner, A novel microfluidic measurements in a magnetic micromixer, Exp. Therm. Fluid Sci. 67 (2015)
mixer-based approach for determining inactivation kinetics of 6–13.
Escherichia coli O157:H7 in chlorine solutions, Food Microbiol. 49 (2015) [253] G.T. Vladisavljević, R.A. Nuumani, S.A. Nabavi, Microfluidic production of
152–160. multiple emulsions, Micromachines 8 (2017) 75.
[228] M.M. Aeinehvand, F. Ibrahim, S.W. Harun, I. Djordjevic, S. Hosseini, H.A. [254] Z.X. Guo, Q. Zeng, M. Zhang, L.Y. Hong, Y.F. Zhao, W. Liu, S.S. Guo, X.Z. Zhao,
Rothan, R. Yusof, Biosensing enhancement of dengue virus using Valve-based microfluidic droplet micromixer and mercury (II) ion detection,
microballoon mixer on centrifugal microfluidic platforms, Biosens. Sens. Actuators A: Phys. 172 (2011) 546–551.
Bioelectron 67 (2015) 424–430. [255] M.C. Giuffrida, L.M. Zanoli, R. D’Agata, A. Finotti, R. Gambari, G. Spoto,
[229] K. Petkovic, G. Metcalfe, H. Chen, Y. Gao, M. Best, D. Lester, Y. Zhu, Rapid Isothermal circular-strand-displacement polymerization of DNA and
detection of Hendra virus antibodies: an integrated device with microRNA in digital microfluidic devices, Anal. Bioanal. Chem. 407 (2015)
nanoparticle assay and chaotic micromixing, Lab. Chip 17 (2017) 169–177. 1533–1543.
[230] J.H. Jung, G.Y. Kim, T.S. Seo, An integrated passive micromixer–magnetic [256] C.J. Ochs, A.R. Abate, Rapid modulation of droplet composition with pincer
separation–capillary electrophoresis microdevice for rapid and multiplex microvalves, Lab. Chip 15 (2015) 52–56.
pathogen detection at the single-cell level, Lab Chip 11 (2011) 3465–3470. [257] B.J. Jin, C. Esteva-Font, A.S. Verkman, Droplet-based microfluidic platform
[231] C. Liu, Y. Li, Y. Li, P. Chen, X. Feng, W. Du, B.F. Liu, Rapid three-dimensional for measurement of rapid erythrocyte water transport, Lab. Chip 15 (2015)
microfluidic mixer for high viscosity solutions to unravel earlier folding 3380–3390.
kinetics of G-quadruplex under molecular crowding conditions, Talanta 149 [258] H. Qu, Y. Yang, Y. Chang, Z. Tang, W. Pang, Y. Wang, H. Zhang, X. Duan,
(2016) 237–243. On-chip integrated multiple microelectromechanical resonators to enable
[232] P. Löb, H. Pennemann, V. Hessel, g/l-Dispersion in interdigital micromixers the local heating, mixing and viscosity sensing for chemical reactions in a
with different mixing chamber geometries, Chem. Eng. J. 101 (2004) 75–85. droplet, Sens. Actuators B: Chem. 248 (2017) 280–287.
[233] P.M. Günther, G.A. Groß, J. Wagner, F. Jahn, J.M. Köhler, Introduction of [259] S. Akbari, T. Pirbodaghi, R.D. Kamm, P.T. Hammond, A versatile microfluidic
surface-modified Au-nanoparticles into the microflow-through device for high throughput production of microparticles and cell
polymerization of styrene, Chem. Eng. J. 135S (2008) S126–S130. microencapsulation, Lab. Chip 17 (2017) 2067–2075.
[234] M.D. Furtaw, D. Lin, L. Wu, J.P. Anderson, Near-infrared metal-enhanced [260] M. Bokharaei, K. Saatchi, U.O. Häfeli, A single microfluidic chip with dual
fluorescence using a liquid–liquid droplet micromixer in a disposable surface properties for protein drug delivery, Int. J. Pharm. 521 (2017) 84–91.
poly(methyl methacrylate) microchip, Plasmonics 4 (2009) 273–280.
[235] K. Matsuyama, K. Mine, H. Kubo, N. Aoki, K. Mae, Optimization methodology
of operation of orifice-shaped micromixer based on micro-jet concept, Biographies
Chem. Eng. Sci. 65 (2010) 5912–5920.
[236] C.H. Yeh, C.H. Chen, Y.C. Lin, Generation of droplets with different
concentrations using gradient-microfluidic droplet generator, Microfluid. Chia-Yen Lee received his BS and MS degrees in Department of Mechanical Engi-
Nanofluid. 11 (2011) 245–253. neering from National Taiwan University, Taiwan in 1991 and 1993, respectively. He
[237] C.H. Yeh, Y.C. Lin, Use of an adjustable microfluidic droplet generator to received his PhD degree in Department of Engineering Science from National Cheng
produce uniform emulsions with different concentrations, J. Micromech. Kung University, Taiwan in 2004. Before he studied for his PhD degree, he worked
Microeng. 23 (2013) 125025. as an engineer, an assistant manager and a section head in TECO, York and DiCon for
[238] R.M. Ribeiro-Costa, M.R. da Cunha, M.R. Gongora-Rubio, P. 3, 1 and 2 years, respectively. He is currently an professor in the Graduate Institute
Michaluart-Júnior, M.I. Ré, Preparation of protein-loaded-PLGA of Materials Engineering at National Pingtung University of Science and Technol-
microspheres by an emulsion/solvent evaporation process employing LTCC ogy. His research interests lie on microfluidics, micro-sensors and applications of
micromixers, Powder Technol. 190 (2009) 107–111. nano-materials to micro-sensors.
[239] K. Matsuyama, K. Mine, H. Kubo, K. Mae, Design of micromixer for
Lung-Ming Fu received M.S. and Ph.D. degrees in Engineering Science from National
emulsification and application to conventional commercial plant for
Cheng Kung University (NCKU), Taiwan, in 1997 and 2001. He had his postdoc
cosmetic, Chem. Eng. J. 167 (2011) 727–733.
training in Department of Engineering Science at NCKU during 2002–2003. He is
[240] A. Montillet, S. Nedjar, M. Tazerout, Continuous production of water-in-oil
currently a distinguished professor in the Graduate Institute of Materials Engi-
emulsion using micromixers, Fuel 106 (2013) 410–416.
neering at National Pingtung University of Science and Technology. His research
[241] S. Maillot, A. Carvalho, J.P. Vola, C. Boudier, Y. Mély, S. Haacke, J. Léonard,
interests are in Microfluidic systems, MEMS fabrication technologies, Micro-sensor
Out-of-equilibrium biomolecular interactions monitored by picosecond
and Computational fluid dynamics.
fluorescence in microfluidic droplets, Lab. Chip 14 (2014) 1767–1774.
[242] J.B. You, K. Kang, T.T. Tran, H. Park, W.R. Hwang, J.M. Kim, S.G. Im,
PDMS-based turbulent microfluidic mixer, Lab. Chip 15 (2015) 1727–1735.

You might also like