You are on page 1of 14

Microfluidics and Nanofluidics (2020) 24:1

https://doi.org/10.1007/s10404-019-2306-y

RESEARCH PAPER

Continuous microfluidic mixing and the highly controlled nanoparticle


synthesis using direct current‑induced thermal buoyancy convection
Kailiang Zhang1 · Yukun Ren1,2 · Likai Hou1 · Ye Tao1 · Weiyu Liu3 · Tianyi Jiang1 · Hongyuan Jiang1,2

Received: 1 August 2019 / Accepted: 12 November 2019 / Published online: 21 November 2019
© Springer-Verlag GmbH Germany, part of Springer Nature 2019

Abstract
We present a flexible and noninvasive approach for efficient continuous micromixing and microreaction based on direct
current-induced thermal buoyancy convection in a single microfluidic unit. Theoretically, microfluids in this microsystem are
unevenly heated by powering the asymmetrically arranged microheater. The thermal buoyancy convection is then formed to
induce microvortices that cause effective fluidic interface disturbance, thereby promoting the diffusion and convective mass
transfer. The temperature distribution and the convection flow in the microchip are first characterized and studied, which can
be flexibly adjusted by changing the DC voltage. Then the mixing performance of the presented method is validated by joint
numerical and experimental analyses. Specifically, at U = 7 V, the mixing efficiencies are higher than 90% as the flow rate
is lower than Qv= 600 nL/s. So high-quality chemical or biochemical reactions needing both suitable heating and efficient
mixing can be achieved using this method. Finally, as one example, we use this method to synthesize nano-sized cuprous
oxide ­(Cu2O) particles by effectively mixing the Benedict’s solution and glucose buffer. Remarkably, the particle size can
be tuned by changing the voltage and the concentration of Benedict’s solution. Therefore, this micromixer can be attractive
for diverse applications needing homogeneous sample mixtures.

Keywords Continuous microfluid mixing · Thermal buoyancy convection · Cuprous oxide particles · Particle synthesis

1 Introduction reactions(Kang et al. 2014; Kulkarni et al. 2017), biological


analyses(Feng et al. 2016) and material syntheses(Britton
Efficient micromixing in continuous flows is a crucial and Raston 2017; Miyakawa et al. 2014). However, continu-
upstream operation prior to synthesis, detecting and analysis ous micromixing is challenging since the fluids in micro-
in various microfluidic-based applications, such as chemical channels are typically in a laminar unmixed state because
of the low Reynolds number and is often dominated by slow
molecular diffusion. For this reason, lots of passive mix-
Electronic supplementary material The online version of this
ing strategies (Lam et al. 2005; Liu et al. 2013; Santana
article (https​://doi.org/10.1007/s1040​4-019-2306-y) contains
supplementary material, which is available to authorized users. et al. 2019; Wang et al. 2018) relying on the special channel
geometry and active ones (Deng et al. 2018; Fu et al. 2013;
* Yukun Ren Pan et al. 2007; Zhang et al. 2018) depending on the exter-
rykhit@hit.edu.cn nal physical fields have been put forward. Concretely, the
* Hongyuan Jiang passive mixers operate without need for additional external
jhy_hit@hit.edu.cn energy but are inflexible as the complicated channel struc-
1
School of Mechatronics Engineering, Harbin Institute tures and long mixing lengths (Cortes-Quiroz et al. 2017;
of Technology, Harbin 150001, Heilongjiang, Yu et al. 2018). The active mixing means, however, usually
People’s Republic of China have simple microchannels and the mixing performance can
2
State Key Laboratory of Robotics and System, Harbin be flexibly manipulated by external fields, such as electric
Institute of Technology, Harbin 150001, Heilongjiang, (Chang and Yang 2009; Chen et al. 2009; Lim et al. 2010;
People’s Republic of China Wu and Li 2007), acoustic(Cui et al. 2016; Lu et al. 2018;
3
School of Electronics and Control Engineering, Wang et al. 2012), and magnetic (Roy et al. 2009; Wang
Chang’an University, Xi’an 710064, Shannxi, et al. 2011).
People’s Republic of China

13
Vol.:(0123456789)
1 Page 2 of 14 Microfluidics and Nanofluidics (2020) 24:1

Recently, thermal-field-based micromixing is becom- efficiency of fluid in the acoustic-thermal approaches is


ing increasingly attractive due to it noninvasive nature and low because of the extra energy consumption for fluid mix-
simple manipulation mechanisms. First, the elevated fluid ing (Qu et al. 2017; Wang et al. 2017). So aiming at these
temperature can highly improve the molecular diffusion pro- three issues, it is meaningful and important to propose
cess since the diffusion coefficient is positively temperature- an efficient, economic and high-throughput thermal-based
dependent (Zhang et al. 2019). More importantly, thermal approach with a low dependence on fluid characteristics
buoyancy convection (Flores–Flores et al. 2015; Zheng et al. for continuous micromixing and microreaction, and to
2018) and thermocapillary convection (Mallea et al. 2017; reveal the mixing mechanism behind it.
Winterer et al. 2018; Yang et al. 2018) can be induced under In this paper, we present a novel method for efficient
a temperature gradient field, which can dramatically dis- continuous micromixing and microreaction based on direct
turb the fluidic interface and then improve the convective current-induced thermal buoyancy convection (shown in
mass transfer process. The thermal buoyancy convection is Fig. 1) in a single microfluidic unit. This microdevice uti-
caused by the inertial effect within the unevenly heated liq- lizes a simple indium tin oxide (ITO)-made asymmetrical
uid media, where the hotter fluid rises up and the cooler one microheater to heat fluids of interest (The electrode and
goes down because the fluid density decreases as the tem- mixing mechanism are shown in Fig. 1b–d), resulting in
perature increases (Kang et al. 2015; Vela et al. 2009). The the formation of thermal buoyancy convection. Convec-
thermocapillary convection originates from the slip of une- tion flow-induced microvortices give rise to effective flu-
venly heated gas–liquid (or liquid–liquid) interfaces (Prad- idic interface disturbance and convective mass transfer.
han and Panigrahi 2015; Velez-Cordero et al. 2014). Lastly, The mixing performance of the presented micromixer is
the elevated temperature can simultaneously trigger the validated by joint numerical and experimental analyses.
chemical reactions requiring heating, such as the synthesis Specifically, at U = 7 V, the mixing efficiencies are higher
of gold nanoparticle (Ftouni et al. 2012) and the polymerase than 90% at a flow rate lower than Qv= 600 nL/s. Finally,
chain reaction (Ahrberg et al. 2016). Therefore, compared to demonstrate this strategy’s capability for simultaneously
with those microsystems utilizing the integration of a mixer conducting reactions, nano-sized cuprous oxide (­ Cu2O)
and a heating unit to trigger the chemical reactions needing microparticles are synthesized by effectively mixing the
heating conditions (Ftouni et al. 2012; Liu et al. 2011), the Benedict’s solution and glucose buffer in this microdevice.
thermal-based strategies are simple and high-efficiency since Remarkably, the particle size can be tuned by changing the
they can enable not only the efficient micromixing but also DC voltage and the concentration of Benedict’s solution.
the simultaneous microreaction using a single unit. Therefore, this microreactor can be attractive for diverse
On the basis of these thermal-induced micromixing applications, especially the biochemical reactions needing
mechanisms, to date, several micromixing devices utiliz- external heat.
ing either the thermal field (Kim et al. 2009; Yesiloz et al.
2017) or the combination of thermal field and other phys-
ical effects (e.g., acoustic and electric fields) (Qu et al.
2017; Zhang et al. 2016) have been proposed. Although
these designs are useful and promising, there are some
practical challenges limiting their further applicability.
First, to our knowledge, the systematic investigation and
demonstration of thermal-induced mixing mechanisms
have been less advanced, and the applicability of these
mixers has not been validated. Second, many parts in
these chips (e.g., the microwave heaters and interdigital
electrodes) are expensive and time-consuming for fabrica-
tion, which definitely limit the popularization, integration
and commercialization. More importantly, these reported
microdevices have limited applied ranges. Specifically,
these presented microchips merely using thermal field are
aimed at either discrete droplets or static liquid media, Fig. 1  a Three-dimensional schematic of the micromixing device. b
resulting in a low throughput regime. As for the strategies Top view of the micromixer. c Side view of the main channel. d Mix-
combining the thermal effect and other physical fields, the ing mechanism: The vortex induced by buoyancy flow can disturb flu-
idic interface and improve the convective mass transfer of two fluids,
thermal-electrical methods have a strict requirement on
the enhanced liquid temperature also enables an improved diffusive
fluid conductivity because of the introduction of electric mass transfer process as the diffusion coefficient D increases with the
field (Ng et al. 2009; Zhang et al. 2016). And the heating temperature. e Picture of the micromixing device

13
Microfluidics and Nanofluidics (2020) 24:1 Page 3 of 14 1

2 Theoretical background The heat transfer in fluids meets the liquid heat transfer
equation,
The fluids in microchannel will be unevenly heated when ( )
𝜕T
the asymmetical microheater is powered. Since the physi- 𝜌0 cp + u ⋅ ∇T − ∇ ⋅ (kf ∇T) = 0 (6)
𝜕t
cal parameters of fluids (e.g., density and viscosity) are all
temperature-dependent, there will be a density difference where, 𝜌0 is the density, cp the heat capacity, kf the thermal
in the unevenly heated microfluids. Then the fluids will conductivity.
be driven by the density difference-induced force (buoy- At the interfaces of different materials, the temperature
ancy force). Finally a three-dimensional convection flow and the normal heat flux are continuous.
is formed in the fluids of interest. The convection flow- Since the microdevice is exposed to the air, the surface-
induced microvortices on cross-sections help to disturb the air satisfies,
fluidic interface and improve the convective mass transfer. ( )
𝜕T
The basic theories describing the heat and mass transfer in −k = h(Ts − Tr ) (7)
𝜕n s
this microdevice are as follows.
where, k is the thermal conductivity, h the convective heat
transfer coefficient, n the external normal line of the outer
2.1 Heat transfer surface of the microchip, Ts the temperature of the chip’s
surface, Tr the temperature of the surrounding air.
The heat in this microdevice comes from the Joule heat-
ing effect of the ITO-made microheater. The temperature
distribution of the microheater meets the heat-conduction 2.2 Mass transfer
differential equation with internal heat source,
The thermal buoyancy convection is governed by the
𝜕T
𝜌I cI = ∇ ⋅ (kI ∇T) + q (1) Navier–Stokes and the continuity equations
𝜕t
where, 𝜌I is the density of ITO film, cI the heat capacity, kI 𝜌0 (u ⋅ ∇)u = 𝜇(T)∇2 u − ∇p + F (8a)
the thermal conductivity, q the Joule heating source.
The Joule heating flux density q is related to the electric ∇⋅u=0 (8b)
current density and electric field intensity,
where, 𝜌0 is the fluid density at room temperature T0 , 𝜇(T)
q = J⋅ E = 𝜎 ⋅ |∇𝜙| 2
(2) the dynamic viscosity at temperature T, p the pressure, F the
buoyancy force, F = −g[𝜌0 − 𝜌(T)] ≈ −g𝜌0 𝛽T (T − T0 )(𝜌(T)
where, q is the Joule heat flux density, J the electric current the fluid density at temperature T, 𝛽T the thermal expansion
density, E the electric field intensity, σ the conductivity of coefficient).
ITO film. 𝜙 the electric potential. At fluid–wall interfaces, the flow field meets the no-slip
The electric potential distribution of the microheater boundary condition: uwall = 0.
meets the Laplace’s equation. The concentration of samples meets the convective-dif-
fusion equation,
∇ ⋅ (𝜎E) = −𝜎∇2 𝜙 = 0 (3)
𝜕c
At the microheater’s surface, the electric current density + (u ⋅ ∇)c = ∇ ⋅ [D(T)∇c] (9)
𝜕t
J meets the following equation,
where, c is the concentration of species, D(T) the diffusion
J⋅n=0 (4) coefficient at the temperature T.
where, n is the unit normal vector. The dimensionless convective-diffusion equation can be
As for the glass substrate and PDMS channel, their tem- given by
perature distributions satisfy the unsteady-state heat-con- 𝜕c∗ 1 2 ∗
duction differential equation without internal heat source, + (u∗ ⋅ ∇)c∗ = ∇c (10)
𝜕t∗ Pe

𝜌gp cgp
𝜕T
= ∇ ⋅ (kgp ∇T) (5) where Pe(T) = UL
D(T)
is the Peclet number, which represents
𝜕t the ratio of convection effect-to-molecular diffusion. U is the
where, 𝜌gp is the density, cgp the heat capacity, kgp the thermal characteristic convection rate, L the characteristic size of
conductivity. microchannel, D(T) the diffusion coefficient at temperature
T.

13
1 Page 4 of 14 Microfluidics and Nanofluidics (2020) 24:1

At the surfaces, inlets and outlet of the microchannel, the also be prepared by the following process. First, 85 grams of
concentration field meets the following boundary conditions, sodium citrate and 50 grams of anhydrous sodium carbonate
are dissolved in 400 mL of deionized water. Then 8.5 grams
⎧ 0 Inlet1 of anhydrous copper sulfate is added to 50 mL of hot water.

⎨ 1 Inlet2 (11) Finally, the Benedict’s solution is obtained by mixing the
⎪ n ⋅ ∇c = 0 Wall,Outlet solutions prepared in the first two steps. The concentration

of glucose buffer used in experiments is 0.2 M. The syn-
where, n is the unit normal vector of channel walls and thesized particles are visualized using a scanning electron
outlet. microscope (SU-8010, Hitachi, Japan).

3.2 Temperature distribution and thermal


buoyancy convection
3 Materials and methods
To analyze the temperature distribution and thermal buoy-
3.1 Device design and sample preparation ancy convection in this microdevice, a three-dimensional
simulation model is first established using a COMSOL soft-
This micromixer is composed of an indium tin oxide (ITO) ware (COMSOL Multiphysics 5.3a). Figure 2a shows an
glass (the thicknesses of the substrate and ITO film are elliptical temperature profile of the substrate–fluid interface
1.1 mm and 1 µm, respectively) in the bottom, a glass slide in the steady state at a DC voltage of 6 V. The temperature
with a height of 140 µm in the middle, and a polydimethyl- is unevenly distributed and dissipated from the microheater
siloxane (PDMS) microchannel in the top. The pattern of the to the surrounding regions. The maximum temperature Tm
microheater is obtained after an etching process. The micro- (marked in Fig. 2a), an important parameter evaluating the
channel is fabricated by pouring and curing the PDMS on a heating characteristic of the microheater, is located at the
polymethylmethlacrylate (PMMA) mold. During the assem- electrode center. Figure 2b exhibits the temperature distribu-
bly process of the microchip, the thin glass slide and the tions of fluids on three cross-sections (marked in Fig. 2a).
ITO glass are first bonded together using UV glue. Then the Specifically, the fluid closing to the microheater has the
thin glass slide and the PDMS channel are assembled after maximum temperature, and the fluid located at the other
an oxygen plasma treatment (ZEPTO, Diener, Germany). side of microchannel has the minimum temperature. Addi-
The dimension parameters of this micromixer are marked in tionally, the temperature distributions on these three cross
Fig. 1b–c. Specifically, w1 = 3 mm, w2 = 1 mm, w3 = 1 mm, sections are slightly different. This phenomenon is explained
w4 = 2 mm, w5 = 0 mm, h1 = 1 μm, h2 = 0.4 mm, h3 = 1 mm, as follows. The generated Joule heat at different positions
and L = 20 mm. In these parameters, the electrode width (w3) of the microheater is same. However, compared with the
and electrode displacement w5 are determined by numeri- electrode element far from the electrode center, the element
cally analyzing their effects on mixing performance (see close to the electrode center is more difficult to dissipate
Fig. 4a–b). The mixing length used in this work is 6 mm. the heat because of the lower temperature difference in its
The fluorescent intensity of Rhodamine B solution is surrounding regions. Therefore, the temperature is unevenly
strongly temperature-dependent. To investigate the temper- distributed along the electrode length. Corresponding to the
ature distribution of fluid, 0.1 mM Rhodamine B solutions temperature profiles on these three cross sections, the flow
are injected into the main channel of the micromixer through fields are given in Fig. 2c. A microvortex crossing the whole
two inlets using a syringe pump (PHD Ultra, Harvard Appa- section is formed when the microheater is powered, which
ratus). A fluorescence microscope (IX73, Olympus, Japan) can significantly disturb the fluidic interface and improve the
with a CMOS camera (Prime 95B, Photometrics) is used convective mass transfer. Because of the slight differences
to visualize the concentration fields and micromixing pro- of temperature distributions on these three cross-sections,
cess. A DC power supply (KXN-305D, Zhao Xin, China) is the induced flow fields are also different from each other.
utilized to power the microheater. The image of the experi- Figure 2d displays the experimental electric current and
mental setup is provided in Fig. S1 of the Supporting Infor- numerical maximum temperature Tm of the microheater as
mation. For visualizing the mixing process of solvents, one a function of DC voltage. It can be seen that the electric
fluid is 0.003 mM fluorescein (Kermel, Tianjin) dissolved current linearly increases with the DC voltage, indicating
in deionized water (or absolute ethanol), the other one is that the resistance of the microheater can be treated as a
deionized water (or absolute ethanol). constant (~ 50 Ω) under different temperatures. The maxi-
The nano-sized ­Cu2O particles are synthesized by mixing mum temperature Tm, however, nonlinearly increases with
the Benedict’s solution (Tianjin Opson Chemical Sales Co., the rising DC voltage. And it can reach 377.2 K as the DC
Ltd, China) and glucose buffer. The Benedict’s solution can voltage is 9 V.

13
Microfluidics and Nanofluidics (2020) 24:1 Page 5 of 14 1

Fig. 2  a–bTemperature distributions on the xy-plane (z = 0 μm) (a) e The temperature difference (Th(x)− Tc(x)) (marked in Fig. 2a) along
and the I–I, II–II, and III–III cross-sections (marked in Fig. 2a) b at the channel length under different voltages when the flow rate Qv
U = 6 V and Qv = 300 nL/s. c Flow fields on the I–I, II–II, and III–III is 300 nL/s. f The average transverse convection velocity along the
cross sections at U = 6 V and Qv = 300 nL/s. (d) The current–voltage channel length under different voltages when the flow rate Qv is 300
and the maximum temperature Tm-voltage curves of the microheater. nL/s

Figure 2e shows the plots of temperature difference strongly temperature-dependent, so the temperature of fluid
Th(x)− Tc(x) (Th(x) and Tc(x) are marked in Fig. 2a) versus chan- can be determined according to the relationship between
nel length L at different DC voltages. It is observed that the local temperature and florescent intensity of the 0.1 mM
Th(x)− Tc(x) decreases along the channel length, whereas, it Rhodamine B solution (Zhao et al. 2013).
can be treated as a constant when L is smaller than 6 mm.
Moreover, Th(x)− Tc(x) increases as the voltage rises, and it T(K) = [A0 + A1 S(T) + A2 S2 (T) + A3 S3 (T)](K) (12)
can reach 11 K at U = 8 V. Figure 2f depicts the relation-
where S(T) = I(T)/IRT is the dimensionless fluorescent inten-
ships between the average transverse convection velocity
sity, I(T) the fluorescent intensity at temperature T, IRT the
(average|uy|) and the channel length under different voltages
fluorescent intensity at room temperature (293.15 K). The
and the fixed flow rate of 300 nL/s. Although the average|uy|
fitting parameters in Eq. (12) are determined by measuring
varies with the channel length, the difference is very small
the fluorescent intensities of 0.1 mM Rhodamine B solution
and can be neglected. And also, the average|uy| is positively
at various temperatures (Fig. 3a). Specifically, A0 = 387.81,
affected by the increased DC voltage. Specifically, the
A1 = −246.58, A2 = 264.17, A3 = −112.25.
average transverse convection velocity increases from 0 to
0.1 mM Rhodamine B solutions are simultaneously
480 μm/s as the DC voltage rise from 0 to 8 V. This can be
injected into the main channel of the micromixers through
explained that the temperature difference positively depends
two separate inlets. The distributions of fluorescent intensity
on the DC voltage (Fig. 2e), and the increased temperature
under different DC voltages are obtained through a fluores-
induces more intensive buoyancy flow (Eq. 8a). Therefore,
cence microscope (IX73, Olympus, Japan) with a CMOS
the buoyancy-flow-induced micromixing process can be con-
camera (Prime 95B) and an image J software. Finally,
trolled by the DC voltage.
the average (Th− Tc) and Tm (Th, Tc, and Tm are marked
To experimentally determine the temperature distribution
in Fig. 2a) can be calculated using Eq. (12). As shown in
of the fluid–substrate interface and validate the accuracy of
Fig. 3b, the experimental temperatures and temperature dif-
numerical analyses, a kind of thermometry based on tem-
ferences match well with the numerical results, proving the
perature-dependent fluorescent intensity is used in this work.
accuracy of above numerical analyses.
That is, the fluorescent intensity of Rhodamine B solution is

13
1 Page 6 of 14 Microfluidics and Nanofluidics (2020) 24:1

To quantitatively evaluate the mixing performance, the


mixing index is defined by the following equation (Feng
et al. 2019; Yesiloz et al. 2017),
� ∑
1∕n (Ii − Iav )2
Me = 1 − (13)
Iav

where, Ii is the concentration (or fluorescent intensity) of


each point, n the total number of the pixels, Iav the average
concentration (or fluorescent intensity). The range of the
mixing index is 0–1. At an unmixed state, the mixing index
is zero. As for the completely mixed sample mixture, the
mixing index is 1. Usually the micromixer has an excellent
mixing performance when the mixing efficiency is higher
than 0.9.
Using this quantitative index, we first investigate the
effects of dimension parameters of the chip (including the
electrode width w3, electrode displacement w5, and the
channel height h3) on mixing performance. According to
Fig. 4 a–b, it is observed that the mixing efficiency is mix-
ing-length-dependent. To quantify the mixing performance
of the chip, the mixing length used in this work is 6 mm.
That is, the concentration distributions used for mixing per-
Fig. 3  a Plot of dimensionless fluorescent intensity versus tempera- formance evaluation are obtained at II–II line (marked in
ture for 0.1 mM Rhodamine B aqueous solution. b Experimental and Fig. 4a). Figure 5a shows the curve of mixing index ver-
numerical results of the average (Th− Tc) and Tm under different volt- sus the electrode width w3 at a fixed flow rate of 300 nL/s
ages and a constant output power of 0.72 W. Specifically, the
mixing efficiency slightly decreases as the electrode width
rises from 0.2 to 1 mm, and then it dramatically declines
when the electrode width is further increased. This variation
4 Results and discussion trend can be explained by the following statements. First, the
increased electrode width will reduce the temperature dif-
4.1 Numerical analysis ference of fluid, leading to a decreased convection velocity.
Besides, a microvortex crossing the whole section changes
After the analysis of thermal buoyancy convection, the into a counter-rotating vortex pair when the electrode width
mass transfer process induced by convection flow is fur- is higher than 1 mm (Fig. 5a). Some previous studies have
ther analyzed. Figure 4a and b compares the mixing per- proved that the microvortex crossing the whole section has
formance of the mixer at different voltages and the fixed the greater disturbance to fluidic interface compared with
flow rate of 900 nL/s. The concentration gradient and the the second vortex pair (Harnett et al. 2008; Zhang et al.
concentration distribution along three cross-sections are 2018). So the changed vortex pair can further decrease the
analyzed and compared. The cross-sections are selected mixing efficiency. Therefore, the mixing efficiency is sig-
at three different positions L 1 = 0 mm, L 2 = 6 mm and nificantly decreased as the electrode width is higher than
L3 = 12 mm along the length of the micromixer. The con- 1 mm. Based on above analyses, when the output power is
centration gradient and concentration distributions at L1, fixed at 0.72 W, although the mixer always has an excellent
L2, and L3 of the mixer are completely different. mixing performance when the electrode width ranges from
When power is off, the concentration gradient contour 0.2 to 1 mm, the microheater needs smaller control voltages
and concentration distribution in main channel show that at w3 = 1 mm. That is, the control signal only varies from 0
the molecular diffusion process cannot mix the microfluids to 8 V at w3 = 1 mm, which makes the presented microfluidic
well (Fig. 4a). While the microheater is powered by a 6-V platform possibly miniaturized by replacing the DC signal
DC voltage, the microfluids are effectively mixed by the generator using small batteries. And also, at w3 = 1 mm, the
transverse convection flow inside the micromixer (Fig. 4b), higher output electric current is easier to measure compared
demonstrating that it is efficient to mix fluids by thermal with results in other smaller electrode width. Therefore, the
buoyancy convection. electrode width of this micromixer is finally set as 1 mm.

13
Microfluidics and Nanofluidics (2020) 24:1 Page 7 of 14 1

Fig. 4  a Simulated concentration distribution when power is off at centration distribution at U = 6 V and Qv= 900 nL/s. b1 exhibits the
Qv= 900 nL/s. a1 exhibits the concentration distribution on xy-plane. concentration distribution on xy-plane. b2–b4 show the concentration
a2–a4 show the concentration distributions on the I–I, II–II, and III– distributions on the I–I, II–II, and III–III section, respectively. b5–b7
III section, respectively. a5–a7 display the concentration gradient display the concentration gradient on I–I, II–II, and III–III section,
on the I–I, II–II, and III–III section, respectively. b Simulated con- respectively

Then we adjust the electrode displacement w5 (marked shown in Fig. 5b, when the DC voltage and flow rate are
in Fig. 1b), the distance between the upper edge of the fixed at 6 V and 300 nL/s, respectively, the mixing efficiency
electrode and the upper side of main channel, to study the decreases from 98.4 to 86.0% as the electrode displacement
influence of electrode position on mixing performance. As w5 increases from 0 to 1 mm. This negative relationship

13
1 Page 8 of 14 Microfluidics and Nanofluidics (2020) 24:1

Fig. 5  a Mixing efficiencies and average transverse convection veloc- ing efficiencies of the final design under different voltages at Qv= 300
ities under different electrode widths at P = 0.72 W and Qv= 300 nL/s. nL/s. e Mixing efficiencies of the final design under different flow
b Mixing efficiencies and average transverse convection velocities rates at U = 6 V. Velocity ratio refers to the ratio of average transverse
under different electrode displacements at U = 6 V and Qv= 300 nL/s. velocity to the average axial velocity. f Mixing efficiencies under dif-
c Mixing efficiencies and average transverse convection velocities ferent viscosities at U = 6 V and Qv= 300 nL/s
under different channel heights at U = 6 V and Qv= 900 nL/s. d Mix-

is due to the fact that the temperature difference decreases on the mixing performance are studied. Figure 5d shows the
and the vortex changes from a vortex crossing the whole mixing efficiencies of the chip under different DC voltages
section to a counter-rotating vortex pair as the electrode dis- at Qv = 300 nL/s. It can be seen that the mixing efficiency
placement rises. Similar to above analyses, the decreased has a positive relationship with the applied voltage. And the
temperature difference and the changed vortex pattern both mixing efficiency can reach as high as 99.1% at U = 8 V and
negatively influence the mixing efficiency. Therefore, the Qv = 300 nL/s. Since the increased flow rate will decrease
final electrode displacement is set as 0 mm based on this the working time and the relative intensity of transverse con-
numerical result. vection flow, the mixing efficiency decreases as the flow rate
The effect of channel height on the mixing efficiency is rises (Fig. 5e). And compared with the mixing efficiency
further analyzed. As shown in Fig. 5c, the mixing efficiency at Qv = 300 nL/s, the result at Qv = 900 nL/s decreases by
is positively influenced by the channel height. Specifically, 12.6%. Finally, we analyze the effect of fluid viscosity on
at the fixed voltage of 6 V and the constant flow rate of mixing performance. As shown in Fig. 5f, the mixing per-
900 nL/s, the mixing efficiency increases by 163.5% (from formance is negatively affected by the fluid viscosity. The
37.3 to 98.3%) as the channel height rises from 0.5 to 3 mm. mixing efficiency decreases from 98.4 to 52.2% when the
This result can be explained that the magnitude of buoy- fluid viscosity rises from 1 to 8 mPa s. This phenomenon
ancy force acting on whole fluid is proportional the cube of is due to the fact that the fluid viscosity will decrease the
the characteristic size of microchannel. And the buoyancy intensity of convection flow. So the convective mass transfer
force positively affects the intensity of thermal buoyancy process can be weakened by the increased fluid viscosity.
convection (Zhang et al. 2019). In previous micromixing
studies, the homogeneous concentration distribution pro- 4.2 Experimental validation
files obtained from the top view of the chip can be used
to represent the excellent mixing state of the whole fluid Besides the numerical analyses, we also perform system-
only when the aspect ratio of microchannel is no less than atic micromixing experiments by injecting deionized (DI)
3 (Chen and Yang 2007; Loucaides et al. 2012; Ng et al. water and fluorescent buffer (fluorescence dissolved in DI
2009). Therefore, the channel width and height are set as water) into the main channel through two separate inlets.
3 mm and 1 mm, respectively. Figure 6a–c shows the concentration distribution profiles
When the dimension parameters of the chip are deter- under different voltages. Specifically, an unmixed laminar
mined, the influences of DC voltage and fluid parameters flow state at Qv = 300 nL/s is observed when power is off

13
Microfluidics and Nanofluidics (2020) 24:1 Page 9 of 14 1

Fig. 6  a–c Concentration distribution profiles under different voltages at Qv= 300 nL/s. d Normalized fluorescent intensities across the channel
width at L(x) = 6 mm

(Fig. 6a). Whereas, the homogeneous mixing of DI water shows the concentration distribution profiles at three flow
and fluorescein is achieved when the microheater is pow- rates and the fixed DC voltage of 6 V. The mixing length
ered. Importantly, compared with the result at U = 4 V used for thorough fluid mixing increases with the rising flow
(shown in Fig. 6b), a shorter mixing length is needed at rates. Figure 7b indicates the plots of mixing efficiency ver-
a higher voltage (Fig. 6c). To quantify the mixing perfor- sus channel length under three flow rates and the fixed volt-
mance, the cross-sectional dye concentration profiles (the age of 6 V. It can be seen that the mixing efficiency increases
dashed lines in Fig. 6) are plotted by measuring the gray with the channel length. This phenomenon can be explained
value of the experimental images. As shown in Fig. 6d, the that the working time of transverse convection flow improves
normalized intensities of the unmixed fluids are 0 for deion- as the channel length rises. Importantly, the mixing perfor-
ized water and 1 for the dyed solution. When the microheater mance can reach 93.4%, a highly efficient mixing state, at
is powered by a 6-V DC voltage, the normalized fluorescent Qv = 300 nL/s and L = 6 mm. Figure 7c exhibits the varia-
intensity is uniform across the microchannel width. tions of mixing efficiency with DC voltage under three flow
According to the quantitative index of mixing efficiency rates and the fixed channel length of 6 mm (See Movie S1
(Eq. 13), we first study the effects of the DC voltage and flow in the Supporting Information). Since the intensity of ther-
rate on the mixing performance of water solution. Figure 7a mal buoyancy convection increases with the rising voltage,

Fig. 7  a Concentration distribution profiles under different flow rates d Mixing efficiencies under different flow rates at U = 6 V and
at U = 6 V. b Mixing efficiencies under different channel lengths at L(x) = 6 mm. e Mixing performances of absolute ethanol under differ-
U = 6 V. c Mixing efficiencies under different voltages at L(x) = 6 mm. ent DC voltages at L(x) = 6 mm

13
1 Page 10 of 14 Microfluidics and Nanofluidics (2020) 24:1

the DC voltage has a positive influence on the mixing effi- During the experimental process, the standard Benedict’s
ciency. Concretely, at a fixed flow rate of 300 nL/s, the mix- solution and glucose buffer with a concentration of 0.2 M are
ing efficiency improves from 13.5 to 97.3% as the voltage simultaneously injected into the main channel. And the volume
ranges from 0 to 8 V. Figure 7 d shows the experimental and flow rate in the main channel is 300 nL/s. Then these two fluids
numerical mixing efficiencies at different flow rates and a are effectively mixed by the transverse convection flow. Mean-
fixed DC voltage of 6 V. The flow rate negatively affects the time, the reaction between the Benedict’s solution and glucose
mixing of water solution and fluorescent dye. Specifically, buffer is triggered by the elevated fluid temperature. Figure 8
the mixing performance decreases from 93.4 to 74.4% as b–c exhibit the SEM images of synthesized C ­ u2O particles
the flow rate rises from 300 to 900 nL/s. This is owing to at U = 5 V (Fig. 8b) and 8 V (Fig. 8c), respectively. It can be
the fact that the working time and the relative intensity of seen that the synthesized ­Cu2O particles are all cubic. The
transverse convection flow decrease with the increased flow particle size distributions at these two different voltages are
rate. Additionally, the experimental and numerical results given in Fig. 8d. Specifically, the average size and the stand-
have a good agreement with each other, demonstrating the ard deviation of particle size at U = 5 V are 0.967 μm and
accuracy of our numerical analyses. 0.12 μm, respectively. However, the synthesized particles are
Besides the mixing of water solution, we further use the much smaller and more uniform at U = 8 V since the average
insulating absolute ethanol as an experimental sample to size and the standard deviation of particle size, respectively,
study the applicability of designed chip on the mixing of decrease to 0.680 μm and 0.09 μm. The influence of DC volt-
insulating fluids. During the experimental process, one solu- age on particle size is further exhibited in Fig. 8e. It is shown
tion is absolute ethanol, the other is fluorescently labeled that the particle size is negatively affected by DC voltage. The
absolute ethanol. Figure 7e shows the variations of mix- particle size decreases by 29.7% as the DC voltage rises from
ing efficiency of absolute ethanol with DC voltage at three 5 to 8 V.
flow rates. It can be seen that the DC voltage also positively Furthermore, we investigate the effect of the concentration
affects the mixing states of ethanol. And the absolute ethanol of Benedict’s solution on particle size. During the experimen-
can reach a highly effective mixing state (Me = 90.8%) at tal process, the DC voltage and the concentration of glucose
U = 6 V and Qv = 300 nL/s. This experiment demonstrates buffer are fixed at 7 V and 0.2 M, respectively. As for the
that the presented design can be used for the mixing of both Benedict’s solution, it is diluted using deionized water, and the
conducting and insulating buffers. Therefore, compared with dilution ratio ranges from 1 to 8. The particle images under dif-
the commonly used electrokinetic micromixing devices rely- ferent dilution ratios are exhibited in Fig. 9a–c. It is observed
ing on the conductivity of buffer solution (Cao et al. 2007; that the dilution ratio negatively affects the particle size.
Jain et al. 2009; Sugioka 2010), the presented method can be According to the statistical result given in Fig. 9d, the particle
used in a broader range of on-chip applications. size declines from 0.793 to 0.477 μm when the dilution ratio
changes from 1 to 8. On the basis of above analyses, the par-
5 Nano‑sized ­Cu2O particle synthesis ticle size can be flexibly tuned by changing the DC voltage
and the concentration of Benedict’s solution. And this particle
Cu2O is a kind of semiconductor material, which has a good synthesis experiment successfully proves the feasibility of the
absorption characteristic to visible light, and the theoretical presented approach in efficient micromixing and simultane-
energy conversion rate can be as high as 14– 20% (Akimoto ous microreaction. Moreover, compared with the synthesis
et al. 2006; Nian et al. 2008). So the ­Cu2O has been widely experiments of C ­ u2O in traditional labs (Dong et al. 2001;
used in solar cells, photo-catalytic degradation of organic pol- Huang et al. 2009), our microfluidic platform is simple and
lutants, photohydrolysis of water for hydrogen, and gas sen- flexible, and the synthesis period is shortened a lot. Therefore,
sors (Kim et al. 2013; Kwon et al. 2015; Lee et al. 2011). this fluid mixing and particle synthesis method is attractive in
To demonstrate the applicability of this thermal microreactor diverse applications including biochemical analysis, material
on material synthesis, ­Cu2O particles are tried to synthesize synthesis, and so on.
by rapidly and efficiently mixing the Benedict’s solution and
glucose buffer. The synthesis mechanism of C ­ u2O particles is
shown in Fig. 8a and Eqs. (14) and (15). 6 Conclusion

Cu2+ + 2OH− → Cu(OH)2 (14) In this paper, we present a novel method for efficient con-
tinuous microfluid mixing and microreaction based on direct
CH2 OH(CHOH)4 CHO + 2Cu(OH)2 current-induced thermal buoyancy convection in a single
microfluidic unit. The fluids of interest are unevenly heated
⟶ ΔCH2 OH(CHOH)4 COOH + 2H2 O + Cu2 O ↓ (15) using an asymmetrically arranged ITO-made microheater,

13
Microfluidics and Nanofluidics (2020) 24:1 Page 11 of 14 1

Fig. 8  a Application of the


presented thermal-based
micromixer for cuprous particle
synthesis. (b–c) SEM images
of the synthesized cuprous
oxide particles at the fixed
glucose buffer concentration
of 0.2 mol/L and different DC
voltages (× 5 k magnification),
the DC voltages in (a–b) are
5 V and 8 V, respectively. d
The particle size distributions
at two different DC voltages. e
The influence of DC voltage on
particle size

leading to the formation of thermal buoyancy convection. this design’s capability for simultaneously conducting reac-
Microfluids then can be effectively mixed by the convection tions, nano-sized C ­ u2O microparticles are synthesized by
flow-induced microvortices. The temperature distribution effectively mixing the Benedict’s solution and glucose buffer
and the natural convection flow in the microchip are first in this microdevice. Remarkably, the particle size can be
studied and characterized, which can be flexibly adjusted by tuned by changing the DC voltage and the concentration of
changing the DC voltage. Then the mixing performance of Benedict’s solution. The particle size decreases from 0.967
the presented micromixer is investigated by joint numerical to 0.680 μm as the DC voltage rises from 5 to 8 V. When the
and experimental analyses. The DC voltage positively affects DC voltage and the concentration of glucose buffer are fixed
the mixing performance. However, the flow rate has a nega- at 7 V and 0.2 M, respectively, the particle size declines
tive effect on mixing efficiency. Specifically, at a fixed flow from 0.793 to 0.477 μm as the dilution ratio of Benedict’s
rate of 300 nL/s, the mixing efficiency improves from 13.5 solution improves from 1 to 8. Therefore, this thermal-based
to 97.3% as the voltage ranges from 0 to 8 V. And the mixing micromixer can be attractive for diverse applications because
performance decreases from 93.4 to 74.4% as the flow rate of its simplicity, high throughput and flexibility.
rises from 300 to 900 nL/s at U = 6 V. Finally, to demonstrate

13
1 Page 12 of 14 Microfluidics and Nanofluidics (2020) 24:1

Fig. 9  (a–c) SEM images of the


­Cu2O particles under different
dilution ratios of Benedict’s
solution (× 5 k magnification).
Scale bar, 2 μm. d Influence of
the dilution ratio of Benedict’s
solution on particle size

Acknowledgements This work is financially supported by the National Cui W et al (2016) Localized ultrahigh frequency acoustic fields
Natural Science Foundation of China (No.11672095, No.11872165, induced micro-vortices for submilliseconds microfluidic mixing.
No. 11702075, No. 11702035). Self-Planned Task (SKLRS201803B) Appl Phys Lett 109:253503. https​://doi.org/10.1063/1.49724​84
of State Key Laboratory of Robotics and System (HIT). Deng Y, Zhou T, Liu Z, Wu Y, Qian S, Korvink JG (2018) Topology
optimization of electrode patterns for electroosmotic micromixer.
Int J Heat Mass Transf 126:1299–1315. https​://doi.org/10.1016/j.
ijhea​tmass​trans​fer.2018.06.065
References Dong Y, Li Y, Wang C, Cui A, Deng Z (2001) Preparation of
Cuprous Oxide Particles of Different Crystallinity. J Colloid
Ahrberg CD, Manz A, Chung BG (2016) Polymerase chain reaction Interface Sci 243:85–89. https​://doi.org/10.1006/jcis.2001.7857
in microfluidic devices. Lab Chip 16:3866–3884. https​://doi. Feng Q, Sun J, Jiang X (2016) Microfluidics-mediated assembly
org/10.1039/c6lc0​0984k​ of functional nanoparticles for cancer-related pharmaceu-
Akimoto K, Ishizuka S, Yanagita M, Nawa Y, Paul GK, Sakurai T tical applications. Nanoscale 8:12430–12443. https​: //doi.
(2006) Thin film deposition of Cu2O and application for solar org/10.1039/c5nr0​7964k​
cells. Sol Energy 80:715–722. https​://doi.org/10.1016/j.solen​ Feng X et al (2019) Effect of vortex on mass transport and mixing
er.2005.10.012 in microcapillary channels. Chem Eng J 362:442–452. https​://
Britton J, Raston CL (2017) Multi-step continuous-flow synthesis. doi.org/10.1016/j.cej.2019.01.055
Chem Soc Rev 46:1250–1271. https​://doi.org/10.1039/c6cs0​ Flores-Flores E et al (2015) Trapping and manipulation of micro-
0830e​ particles using laser-induced convection currents and photo-
Cao J, Cheng P, Hong FJ (2007) A numerical study of an electrother- phoresis. Biomed Optics Express 6:4079–4087. https​: //doi.
mal vortex enhanced micromixer. Microfluid Nanofluid 5:13–21. org/10.1364/BOE.6.00407​9
https​://doi.org/10.1007/s1040​4-007-0201-4 Ftouni J, Penhoat M, Addad A, Payen E, Rolando C, Girardon JS
Chang C-C, Yang R-J (2009) Chaotic mixing in electro-osmotic flows (2012) Highly controlled synthesis of nanometric gold parti-
driven by spatiotemporal surface charge modulation. Phys Fluids cles by citrate reduction using the short mixing, heating and
21:052004. https​://doi.org/10.1063/1.31391​62 quenching times achievable in a microfluidic device. Nanoscale
Chen JK, Yang RJ (2007) Electroosmotic flow mixing in zigzag micro- 4:4450–4454. https​://doi.org/10.1039/c2nr1​1666a​
channels. Electrophoresis 28:975–983. https​://doi.org/10.1002/ Fu L-M, Ju W-J, Tsai C-H, Hou H-H, Yang R-J, Wang Y-N (2013)
elps.20060​0470 Chaotic vortex micromixer utilizing gas pressure driving force.
Chen JK, Weng CN, Yang RJ (2009) Assessment of three AC electroos- Chem Eng J 214:1–7. https​://doi.org/10.1016/j.cej.2012.10.032
motic flow protocols for mixing in microfluidic channel. Lab Chip Harnett CK, Templeton J, Dunphy-Guzman KA, Senousy YM,
9:1267–1273. https​://doi.org/10.1039/b8195​47a Kanouff MP (2008) Model based design of a microfluidic mixer
Cortes-Quiroz CA, Azarbadegan A, Zangeneh M (2017) Effect of chan- driven by induced charge electroosmosis. Lab Chip 8:565–572.
nel aspect ratio of 3-D T-mixer on flow patterns and convective https​://doi.org/10.1039/b7174​16k
mixing for a wide range of Reynolds number. Sens Actuators B Huang L, Peng F, Yu H, Wang H (2009) Preparation of cuprous
239:1153–1176. https​://doi.org/10.1016/j.snb.2016.08.116 oxides with different sizes and their behaviors of adsorption,
visible-light driven photocatalysis and photocorrosion. Solid

13
Microfluidics and Nanofluidics (2020) 24:1 Page 13 of 14 1

State Sci 11:129–138. https​://doi.org/10.1016/j.solid​state​scien​ illumination. International J Hydrog Energy 33:2897–2903. https​
ces.2008.04.013 ://doi.org/10.1016/j.ijhyd​ene.2008.03.052
Jain M, Yeung A, Nandakumar K (2009) Induced charge electro Pan Y-J, Ren C-M, Yang R-J (2007) Electrokinetic flow focusing and
osmotic mixer: Obstacle shape optimization. Biomicrofluidics valveless switching integrated with electrokinetic instability for
3:22413. https​://doi.org/10.1063/1.31672​79 mixing enhancement. J Micromech Microeng 17:820–827. https​
Kang HW, Leem J, Yoon SY, Sung HJ (2014) Continuous synthesis ://doi.org/10.1088/0960-1317/17/4/020
of zinc oxide nanoparticles in a microfluidic system for photovol- Pradhan TK, Panigrahi PK (2015) Thermocapillary convection
taic application. Nanoscale 6:2840–2846. https:​ //doi.org/10.1039/ inside a stationary sessile water droplet on a horizontal surface
c3nr0​6141h​ with an imposed temperature gradient. Exp Fluids. https​://doi.
Kang Z, Chen J, Wu SY, Chen K, Kong SK, Yong KT, Ho HP (2015) org/10.1007/s0034​8-015-2051-2
Trapping and assembling of particles and live cells on large-scale Qu H et al (2017) On-chip integrated multiple microelectromechanical
random gold nano-island substrates. Sci Rep 5:9978. https​://doi. resonators to enable the local heating, mixing and viscosity sens-
org/10.1038/srep0​9978 ing for chemical reactions in a droplet. Sens Actuators B: Chem
Kim SJ, Wang F, Burns MA, Kurabayashi K (2009) Temperature-pro- 248:280–287. https​://doi.org/10.1016/j.snb.2017.03.173
grammed natural convection for micromixing and biochemical Roy T, Sinha A, Chakraborty S, Ganguly R, Puri IK (2009) Mag-
reaction in a single microfluidic chamber. Anal Chem 81:4510– netic microsphere-based mixers for microdroplets. Phys Fluids
4516. https​://doi.org/10.1021/ac900​512x 21:027101. https​://doi.org/10.1063/1.30726​02
Kim J, Kwon Y, Lee H (2013) Metal ion-assisted reshaping of Cu2O Santana HS, Silva JL, Taranto OP (2019) Optimization of micromixer
nanocrystals for catalytic applications. J Mater Chem A 1:14183. with triangular baffles for chemical process in millidevices. Sens
https​://doi.org/10.1039/c3ta1​3182c​ Actuators B: Chem 281:191–203. https​: //doi.org/10.1016/j.
Kulkarni JA, Tam YYC, Chen S, Tam YK, Zaifman J, Cullis PR, Bis- snb.2018.10.089
was S (2017) Rapid synthesis of lipid nanoparticles containing Sugioka H (2010) Chaotic mixer using electro-osmosis at finite Peclet
hydrophobic inorganic nanoparticles. Nanoscale 9:13600–13609. number. Phys Rev E 81:036306. https​://doi.org/10.1103/PhysR​
https​://doi.org/10.1039/c7nr0​3272b​ evE.81.03630​6
Kwon Y, Soon A, Han H, Lee H (2015) Shape effects of cuprous oxide Vela E, Hafez M, Régnier S (2009) Laser-induced thermocapillary
particles on stability in water and photocatalytic water splitting. convection for mesoscale manipulation. Int J Optomechatronics
J Mater Chem A 3:156–162. https​://doi.org/10.1039/c4ta0​4863f​ 3:289–302. https​://doi.org/10.1080/15599​61090​33894​77
Lam YC, Chen X, Yang C (2005) Depthwise averaging approach to Velez-Cordero JR, Velazquez-Benitez AM, Hernandez-Cordero J
cross-stream mixing in a pressure-driven microchannel flow. (2014) Thermocapillary flow in glass tubes coated with pho-
Microfluid Nanofluid 1:218–226. https​://doi.org/10.1007/s1040​ toresponsive layers. Langmuirs 30:5326–5336. https ​ : //doi.
4-004-0013-8 org/10.1021/la404​221p
Lee S, Liang CW, Martin LW (2011) Synthesis, control, and characteri- Wang S, Huang X, Yang C (2011) Mixing enhancement for high vis-
zation of surface properties of Cu(2)O nanostructures. ACS Nano cous fluids in a microfluidic chamber. Lab Chip 11:2081–2087.
5:3736–3743. https​://doi.org/10.1021/nn200​1933 https​://doi.org/10.1039/c0lc0​0695e​
Lim CY, Lam YC, Yang C (2010) Mixing enhancement in microfluidic Wang S, Huang X, Yang C (2012) Microfluidic bubble genera-
channel with a constriction under periodic electro-osmotic flow. tion by acoustic field for mixing enhancement. J Heat Transfer
Biomicrofluidics 4:14101. https​://doi.org/10.1063/1.32797​90 134:051014. https​://doi.org/10.1115/1.40057​05
Liu Z, Lu Y, Yang B, Luo G (2011) Controllable Preparation of Wang Z, Zhang H, Yang Y, Qu H, Han Z, Pang W, Duan X (2017)
Poly(butyl acrylate) by Suspension Polymerization in a Coaxial Wireless Controlled Local Heating and Mixing Multiple Drop-
Capillary Microreactor. Ind Eng Chem Res 50:11853–11862. lets Using Micro-Fabricated Resonator Array for Micro-Reac-
https​://doi.org/10.1021/ie201​497b tor Applications. IEEE Access 5:25987–25992. https​: //doi.
Liu Y, Deng Y, Zhang P, Liu Z, Wu Y (2013) Experimental investiga- org/10.1109/acces​s.2017.27662​70
tion of passive micromixers conceptual design using the layout Wang H, Shi L, Zhou T, Xu C, Deng Y (2018) A novel passive micro-
optimization method. J Micromech Microeng 23:075002. https​:// mixer with modified asymmetric lateral wall structures. Asia-Pac
doi.org/10.1088/0960-1317/23/7/07500​2 J Chem Eng 13:e2202. https​://doi.org/10.1002/apj.2202
Loucaides N, Ramos A, Georghiou GE (2012) Configurable AC elec- Winterer F, Maier CM, Pernpeintner C, Lohmuller T (2018) Optoflu-
troosmotic pumping and mixing. Microelectron Eng 90:47–50. idic transport and manipulation of plasmonic nanoparticles by
https​://doi.org/10.1016/j.mee.2011.04.007 thermocapillary convection. Soft Matter 14:628–634. https​://doi.
Lu Y et al (2018) On-chip acoustic mixer integration of electro-micro- org/10.1039/c7sm0​1863k​
fluidics towards in situ and efficient mixing in droplets. Micro- Wu Z, Li D (2007) Mixing and flow regulating by induced-charge
fluidics Nanofluidics. https:​ //doi.org/10.1007/s10404​ -018-2169-7 electrokinetic flow in a microchannel with a pair of conduct-
Mallea RT, Bolopion A, Beugnot J-C, Lambert P, Gauthier M (2017) ing triangle hurdles. Microfluid Nanofluid 5:65–76. https​://doi.
Laser-Induced Thermocapillary Convective Flows: a New org/10.1007/s1040​4-007-0227-7
Approach for Noncontact Actuation at Microscale at the Fluid/ Yang J et al (2018) Size-tunable capture of mesoscopic matters using
Gas Interface. IEEE/ASME Trans Mechatron 22:693–704. https​ thermocapillary vortex. Appl Phys Lett 113:131602. https​://doi.
://doi.org/10.1109/tmech​.2016.26398​21 org/10.1063/1.50378​62
Miyakawa M, Hiyoshi N, Nishioka M, Koda H, Sato K, Miyazawa A, Yesiloz G, Boybay MS, Ren CL (2017) Effective thermo-capillary mix-
Suzuki TM (2014) Continuous syntheses of Pd@Pt and Cu@Ag ing in droplet microfluidics integrated with a microwave heater.
core-shell nanoparticles using microwave-assisted core particle Anal Chem 89:1978–1984. https​://doi.org/10.1021/acs.analc​
formation coupled with galvanic metal displacement. Nanoscale hem.6b045​20
6:8720–8725. https​://doi.org/10.1039/c4nr0​0118d​ Yu N et al (2018) Integrated obstacle microstructures for gas-liquid
Ng WY, Goh S, Lam YC, Yang C, Rodriguez I (2009) DC-biased separation and flow switching in microfluidic networks. Sens
AC-electroosmotic and AC-electrothermal flow mixing in micro- Actuators B: Chem 256:735–743. https​: //doi.org/10.1016/j.
channels. Lab Chip 9:802–809. https:​ //doi.org/10.1039/b81363​ 9d snb.2017.09.207
Nian J-N, Hu C-C, Teng H (2008) Electrodeposited p-type Cu2O for
H2 evolution from photoelectrolysis of water under visible light

13
1 Page 14 of 14 Microfluidics and Nanofluidics (2020) 24:1

Zhang F, Chen H, Chen B, Wu J (2016) Alternating current electrother- Zhao Y, Zhao C, He J, Zhou Y, Yang C (2013) Collective effects
mal micromixer with thin film resistive heaters. Adv Mech Eng on thermophoresis of colloids: a microfluidic study within the
8:168781401664626. https:​ //doi.org/10.1177/168781​ 40166​ 46264​ framework of DLVO theory. Soft Matter 9:7726. https​://doi.
Zhang K, Ren Y, Hou L, Feng X, Chen X, Jiang H (2018) An efficient org/10.1039/c3sm2​7720h​
micromixer actuated by induced-charge electroosmosis using Zheng J, Xing X, Yang J, Shi K, He S (2018) Hybrid optofluidics
asymmetrical floating electrodes. Microfluidics Nanofluidics and three-dimensional manipulation based on hybrid photo-
22:10–12. https​://doi.org/10.1007/s1040​4-018-2153-2 thermal waveguides. NPG Asia Mater 10:340–351. https​://doi.
Zhang K, Ren Y, Tao Y, Liu W, Jiang T, Jiang H (2019) Efficient micro/ org/10.1038/s4142​7-018-0026-5
nanoparticle concentration using direct current-induced thermal
buoyancy convection for multiple liquid media. Anal Chem Publisher’s Note Springer Nature remains neutral with regard to
91:4457–4465. https​://doi.org/10.1021/acs.analc​hem.8b051​05 jurisdictional claims in published maps and institutional affiliations.

13

You might also like