You are on page 1of 8

Flow characterization in converging-

diverging microchannels
Cite as: Phys. Fluids 30, 112004 (2018); https://doi.org/10.1063/1.5048322
Submitted: 12 July 2018 • Accepted: 23 October 2018 • Published Online: 08 November 2018

Ran Tao, Yakang Jin, Xiang Gao, et al.

ARTICLES YOU MAY BE INTERESTED IN

Oil-water displacements in rough microchannels


Physics of Fluids 30, 112101 (2018); https://doi.org/10.1063/1.5053625

“Phase diagram” for viscoelastic Poiseuille flow over a wavy surface


Physics of Fluids 30, 113101 (2018); https://doi.org/10.1063/1.5057392

Numerical investigation of electrowetting-based droplet splitting in closed digital


microfluidic system: Dynamics, mode, and satellite droplet
Physics of Fluids 30, 112001 (2018); https://doi.org/10.1063/1.5049511

Phys. Fluids 30, 112004 (2018); https://doi.org/10.1063/1.5048322 30, 112004

© 2018 Author(s).
PHYSICS OF FLUIDS 30, 112004 (2018)

Flow characterization in converging-diverging microchannels


Ran Tao, Yakang Jin, Xiang Gao, and Zhigang Lia)
Department of Mechanical and Aerospace Engineering, The Hong Kong University of Science and Technology,
Clear Water Bay, Kowloon, Hong Kong
(Received 12 July 2018; accepted 23 October 2018; published online 8 November 2018)

Experiments are conducted to investigate fluid flows in converging-diverging microchannels


(CDMCs). A new dimensionless number related to channel geometry, Gm, is introduced to com-
bine with the Reynolds number, Re, to characterize the flows. It is found that the new dimensionless
number, ReG = Re · Gm, is more appropriate than Re for flow characterization in CDMCs. Flows are
laminar for ReG < 40 regardless of the geometry of CDMCs. For laminar flows, the flow resistance
model developed in the literature works well. For transitional and turbulent flows, a general scaling law
for the flow resistance is developed, which suggests a polynomial dependence of pressure drop on the
flow rate. Numerical simulations have also been performed to confirm experimental results. Published
by AIP Publishing. https://doi.org/10.1063/1.5048322

INTRODUCTION concerned with the laminar regime, especially for the studies
of flow resistance.27–30 In the laminar regime, the flow resis-
Due to tremendous applications in a variety of areas,
tance has been found to be independent of Re. For turbulent
microscale fluid flows have been extensively investigated. At
flows, how the flow resistance is affected by various parameters
the early stage of microfluidics studies, most attention was
is unknown.
focused on fluid flows in uniform channels,1,2 for which the
In this work, we conduct experiments and numerical sim-
boundary conditions (e.g., velocity slip) and flow scenarios
ulations to investigate flows in CDMCs with rectangular cross
were examined. Compared with uniform channel flows, flows
sections and smooth surfaces. A new dimensionless num-
in non-uniform microchannels are less studied. Converging-
ber, Gm, which is associated with the geometry of CDMCs,
diverging microchannels (CDMCs) are representative non-
is proposed. It is found that the product of Re and Gm,
uniform channels and the flows in CDMCs also have great
ReG = Re · Gm, forms an appropriate dimensionless number
applications in biology, medicine, and engineering, such as
for the characterization of flow transitions in CDMCs. In addi-
DNA stretching,3 artery stenosis (narrowing of an artery),4–6
tion, the flow resistance for both laminar and turbulent flows
particle separation,7,8 micromixers,9–12 design of microflu-
is investigated. For laminar flows, the theoretical prediction
idic rheometry,13,14 flow rectifications,15 and heat transfer
developed in the literature works well. For turbulent flows, a
enhancement,16 just to name a few. These applications require
general law is obtained for describing the dependence of flow
a deep, fundamental understanding of flow characteristics in
resistance on ReG , which suggests a polynomial relationship
CDMCs.
between the pressure drop and the flow rate.
In the literature, many attempts have been made to investi-
gate flows in macroscale converging-diverging channels,17–26
CHIP FABRICATION AND EXPERIMENTAL SETUP
for which pressure/velocity distributions, and flow scenarios
have been studied. In the previous work, the flow dynamics in The traditional photolithography method31–33 was
converging-diverging channels is mainly characterized by the employed to fabricate the microchannels in an n-type sili-
Reynolds number, Re. For uniform channel flows, it is well- con wafer. Deep reactive ion etching (DRIE) (Surface Tech-
known that Re is an important number for flow description. In nology systems, United Kingdom) was used to construct
converging-diverging channels, however, Re may not be suf- converging-diverging microchannels with an etching rate
ficient to describe the flow dynamics. For instance, it is found of 2.5 µm/min. Fluid reservoirs were defined and etched
that flows in a mild converging-diverging tube are laminar for through DRIE from the backside of the wafer. To ensure
Re < 900,17 while flows in converging-diverging channels with uniform properties of the channel surface, wet oxidation of
a relatively large converging/diverging angle show instability the wafer was performed to generate a 300-nm-thick oxide
and even turbulence when Re > 400.6 In other converging- layer before the microchannels were enclosed by a 500-µm-
diverging channels, flows remain laminar for Re ranging from thick glass through anodic bonding. Before bonding, channels
100 to 1000.18 In CDMCs, a similar problem is expected. were investigated by a scanning electron microscope (SEM)
Unfortunately, little work has been conducted to characterize (JEOL JSM-6490). The channel height was determined by a
flows in CDMCs,27–30 which require extensive investigation. surface profiler (Tencor P-10, ClassOne Equipment, USA).
Furthermore, most of the previous work on CDMC flows is Microchannels of six different geometries were fabricated.
All of them were symmetric converging-diverging channels
with a linearly varying cross-sectional area, as illustrated
a) Email: mezli@ust.hk Fig. 1(a). Figure 1(b) shows an SEM image of a CDMC.

1070-6631/2018/30(11)/112004/7/$30.00 30, 112004-1 Published by AIP Publishing.


112004-2 Tao et al. Phys. Fluids 30, 112004 (2018)

FIG. 2. Experimental setup of the fluidic system. Arrows indicate the flow
direction.

Swagelok, USA) were used to measure the pressures at the


inlet and outlet reservoirs through a data acquisition module
(NI9203 and NI cDAQ9171) installed with a labview pro-
gram (LABVIEW 2013, National Instruments, USA). For each
channel geometry, two microfluidic devices were fabricated
and the results were found to be similar. For each flow, at least
three experiments were conducted and the average results are
presented.
To explore the flow details, some flows were visual-
ized through streak photolithography, for which DI water was
seeded with fluorescent tracers and polymer microspheres with
a diameter of ∼3.2 µm (R0300, Duke Scientific, USA) (the
maximum Stokes number, St = 0.14, occurs at the throat of
the channels, which is quite small and indicates that the tracer
particles could follow the flow streamlines well). The flow
FIG. 1. Illustration of non-uniform microchannels. (a) Schematic of a behavior was recorded by an inverted epifluorescence micro-
converging-diverging microchannel. (b) SEM image of a typical converging-
scope (Eclipse Ti-U, Nikon, Japan) equipped with a CCD
diverging microchannel.
camera (EXi blue, QImaging, Canada).

The geometric parameters for the microchannels are listed RESULTS AND DISCUSSION
in Table I. By varying the flow rate, the upstream and downstream
The device was washed in an acetone ultrasonic bath for pressures, P1 and P2 , were measured. Figure 3 plots the
three times and then rinsed in deionized (DI) water. After
that, the fluidic chip was packaged by a PMMA holder. The
schematic of the experimental setup is illustrated in Fig. 2,
similar to our previous work.32,33 Before each experiment,
the chip was examined under a microscope to ensure that
there was no residue inside the channel and DI water was
prefilled into the channel to avoid air trapping. In the exper-
iments, DI water with a specific flow rate was achieved by
a stainless steel syringe (Cole-Parmer Instrument, USA) and
a high performance syringe pump (Nexus 6000, Chemyx,
Inc., USA). Pressure transducers (25 MPa with an accuracy
of 0.5% limit point calibration and 0.25% best-fit straight line,

TABLE I. Geometric parameters of six microchannels.

Microchannel I II III IV V VI

Minimum width wmin (µm) 8.65 16.31 5.91 10.75 16.56 10.59
Maximum width wmax (µm) 532 540 880 885 891 1773
FIG. 3. Pressure drop as a function of flow rate. Solid and open symbols are
Height h (µm) 60 60 60 60 60 60 the current and previous experimental results. Solid lines represent Eq. (12)
Length l (mm) 10 10 10 10 10 10 in the transitional and turbulent regimes. Dashed lines are the predictions of
Taper angle θ (◦ ) 3 3 5 5 5 10 Eq. (4) for the laminar regime. The inset shows the data in a linear plot. Error
bars are not shown as they are less than 3%.
112004-3 Tao et al. Phys. Fluids 30, 112004 (2018)

pressure drop over the CDMCs, ∆p = P1 − P2 , as a func- For the previous results,28 Re∗ should be higher than 100. The
tion of the flow rate Q, where wmin and θ are the channel width dependence of Re∗ on the channel geometry suggests that Re
at the throat and the taper angle of the channel, respectively, might not be an appropriate quantity for the characterization
as depicted in Fig. 1(a). It is seen that the ∆p-Q curves expe- of flows in CDMCs. Actually, the hydraulic diameter Dh in
rience a transition from a linear relationship to a nonlinear Re only gives the average size of the channel but does not
dependence as the flow rate is increased. The transition indi- reveal the converging/diverging rate. To consider the effects of
cates that the flows might undergo changes in flow scenarios channel geometry on the flow resistance, a geometry-related
as Q is varied. At low flow rates, the flows should be in the dimensionless number is proposed, which is given by
laminar regime and the linear relationship between Q and ∆p is r
wmax
consistent with previous theoretical and experimental results Gm = tan θ, (1)
(Fig. 3).27,28 At high flow rates, the flows might turn to turbu- wmin
lence and the quantitative nonlinear dependence of ∆p on Q where θ is the taper angle of the channel. In Eq. (1), θ describes

has not been studied before. the converging/diverging rate of the channel and wmax wmin
For non-uniform channels, the averaged channel size is related to the channel length for a specific θ. A large θ
is usually employed for flow characterization.34,35 For the

or wmax wmin leads to a strong flow velocity variation in the
linearly varying cross section microchannels in this work, channel and enhances the flow instability. Therefore, Gm is a
the averaged hydraulic diameter Dh = 4A/λ appears to be good number accounting for the effect of channel geometry
a proper parameter, where A and λ are the cross-sectional on flows, especially the flow transition. Then, the product of
area and the perimeter at the middle of the converging (or Re and Gm, which is denoted by ReG ,
diverging) part of the channel. Then, Re can be expressed as
ρV Dh wmax
r
Re = ρVDh /µ, where ρ and µ are the fluid density and vis- ReG = Re · Gm = tan θ, (2)
cosity, respectively, and V is the mean velocity at the middle µ wmin
point of the converging part. Figure 4 depicts the flow resis- forms another dimensionless number that contains both the
tance, Rf = ∆p/Q, as a function of Re, where previous exper- geometric influence and the competition between viscous and
imental results for similar converging-diverging channels are inertial effects, which could be a suitable number for char-
also shown. For each channel geometry in the current exper- acterizing flow transition. It is noted that the square root in
iments, it is seen that there exists a critical Re, Re∗ , where Eq. (1) is used to ensure that Re plays a dominant role over
flow transition takes place. For Re < Re∗ , Rf is independent the geometry effect in flow description.
of Re. However, Rf increases greatly with increasing Re when Figure 5 shows Rf as a function of ReG . It is seen that
Re > Re∗ . For previous experiments, Rf remains constant for flow transitions take place at the same ReG value, ReG ≈ 40,
Re up to 100. regardless of the channel geometry. This is consistent with
It is known that Rf is independent of Re for laminar flows. the flow observation shown in Fig. 6. The streakline patterns
Therefore, Fig. 4 indicates that previous experiments27,28 are in Fig. 6 demonstrate that flows are laminar for ReG < 40
in the laminar regime and the flows in the current experiments in channels with different geometries (channels IV and VI in
cover both laminar and turbulent regimes. Furthermore, Fig. 4 Table I; a video for the flow in channel IV can be found in
shows that the critical Re, Re∗ , depends on the geometry of the supplementary material). The flows appear to be transi-
the channel, which is consistent with previous work.6,17–20 tional for 40 < ReG < 220 and become fully turbulent when
For the six channels in Table I, channels I, II, III, IV, V, ReG > 220. For previous experiments, ReG < 40, and therefore,
and VI, Re∗ is about 20, 30, 10, 15, 20 and 8, respectively. the flows are in the laminar regime and Rf remains constant, as

FIG. 4. Flow resistance Rf versus Re. FIG. 5. Flow resistance Rf as a function of ReG .
112004-4 Tao et al. Phys. Fluids 30, 112004 (2018)

FIG. 6. Streaklines of various flows. The red lines show


the position of the channel walls. [(a)–(f)] Channel IV in
Table I. [(g)–(l)] Channel VI in Table I. (a) Q = 0.05 ml/h,
ReG = 0.17. (b) Q = 2 ml/h, ReG = 6.6. (c) Q = 20 ml/h,
ReG = 66.4. (d) Q = 60 ml/h, ReG = 199.1. (e) Q = 70
ml/h, ReG = 232.3. (f ) Q = 80 ml/h, ReG = 265.5. (g) Q
= 2 ml/h, ReG = 7.2. (h) Q = 10 ml/h, ReG = 35.9. (i) Q
= 20 ml/h, ReG = 71.7. ( j) Q = 50 ml/h, ReG = 179.3. (k)
Q = 70 ml/h, ReG = 251.0. (l) Q = 80 ml/h, ReG = 286.8.

shown in Fig. 5. A comparison between Figs. 4 and 5 indicates to dominate. In this case, the total flow resistance can be con-
that ReG is more appropriate than Re for flow characterization sidered as a combination of the resistances caused by viscous
in CDMCs. (Rl ) and inertial (Rin ) effects such that
In CDMCs, the channel width w varies in the flow direc-
Rf = Rl + Rin . (5)
tion. For laminar flows, the flow resistance depends on the ratio
of w to the channel height h. At the locations, where w  h or If Eq. (5) is divided by Rl , the reduced flow resistance,
w  h, the Poiseuille’s law works well. For w ∼ h, flow resis- Rf∗ = Rf Rl , is written as

tance needs special treatment, which has been suggested in the
literature.27,36 Then, the flow resistance for laminar flows in Rf∗ = 1 + Rin

, (6)
CDMCs, Rl , can be estimated as where Rin ∗ = R  R . Figure 7 shows R∗ as a function of Re
in l G
  x2 ∗ in
in the transitional and turbulent regimes (ReG > 40). It is seen
12µ x1 Ip
"
dx
Rl = 2 + 16π µ
2
dx ∗ is sensitive to the taper angle. To obtain a universal
that Rin
h 0 (wmin + 2 tan θx)3 x1 A(x)
2
∗ on Re , the effect of the taper
 law about the dependence of Rin G
12µ L/2
#
dx angle needs to be considered. For this purpose, a dimensionless
+ 3 , (3)
h x2 (wmin + 2 tan θx) quantity, S m , is proposed, which is given by
where x = 0 is at the throat of the channel, x 1 and x 2 are the
locations, where w/h = 1/4 and w/h = 4, respectively, A(x) is the
cross-sectional area, and Ip∗ = Ip A2 with Ip = ∫ A (y2 + z2 )dA,

which is called the specific polar moment of cross-sectional
inertia.27,36 For channels
 of rectangular
 cross section, it has
been shown that Ip∗ = w2 + h2 (12wh).37 Integrating Eq. (3),
it is obtained that
!
6µ 1 16 424µ 12µ wmax
Rl = − 2 + 3 + 3 ln .
h tan θ wmin 2 h 3h tan θ h tan θ 4h
(4)

The predictions of Eq. (4) are depicted as the dashed lines


in Fig. 3, which are generally in good agreement with
the experiments. The discrepancy for the smallest channel
with wmin = 5.91 µm might be caused by the error in the
determination of wmin and the fabrication defects at the
throat.
In the laminar regime, the flow resistance is mainly caused
by the viscous effect. When ReG > 40, the inertial effect tends ∗ versus Re in the transitional and turbulent regimes.
FIG. 7. Rin G
112004-5 Tao et al. Phys. Fluids 30, 112004 (2018)

h Therefore, the ∆p–Q relationship for flows in CDMCs for


Sm = . (7)
λ any ReG value (or in any regime) can be obtained using either
It is noted that S m can also be expressed using the taper angle Eq. (4) or Eq. (12). These two expressions indicate that ReG
θ as S m = h/(2h + l tan θ + 2wmin ). For a given wmin , at a is indeed an appropriate quantity for flow characterization in
relatively large θ, Rin∗ is large, as shown in Fig. 7, while S is
m CDMCs. The physical meaning of ReG suggests that a large
small. Therefore, the dependence of Rin ∗ S on Re for different
m G Gm promotes flow transition, i.e., flows can change from lam-
θ could be unified. inar to turbulent at a relatively small Re (or velocity) if the
∗ · S as a function of Re . For Re > 40,
Figure 8 plots Rin Gm of a CDMC is large. According to Eq. (1), a large Gm
m G G
∗ · S for different geometries col-
it is seen that the values of Rin m can be achieved by using a large taper angle θ. This leads to a
lapse into a single linear relationship, which can be expressed large velocity variation transverse to the flow direction, which
as enhances the flow instability and consequently tends to gen-

Rin Sm = k(ReG − 40) for ReG > 40, (8) erate turbulence. Another reason accounting for this effect is
that a large θ can cause flow detachment from the channel sur-
where k = 3.14 × 10−4 is the slope of the dashed line in Fig. 8, face in the divergent part due to the inertial effect. As shown
which is independent of the geometry of the channels. On the in Figs. 6(c) and 6(i), the fluid could not follow the channel
basis of Eq. (8), it is easy to obtain that surface when ReG becomes higher than 40. Without the con-
λk finement of the channel surface, the flow stability is weakened.

Rin = (ReG − 40). (9) When the flow velocity is further increased, turbulence can be
h
∗ can be written as
easily triggered.
If the flow rate Q = VA is used, Rin
4k ρGm 40kλ NUMERICAL SIMULATION

Rin = Q− , (10)
µh h Numerical simulations have also been performed to con-
and the flow resistance Rf is determined as firm the experiments. As the Navier-Stokes equation is valid
for microchannel flows,39 the commercial software FLUENT
4k ρGmRl
!
40kλ is used. The structure of channel VI in Table I is employed.
Rf = Q+ 1− Rl , (11)
µh h Laminar and k-epsilon models are used for laminar and tur-
bulent flows, respectively. Figure 9 shows the dependence of
which leads to
∆p on Q as Q is varied. It is seen that numerical results are in
∆p = aQ + bQ2 for ReG > 40, (12) good agreement with experiments over the flow rate range in
the experiments.
Rl and b = 4kρGmR
 
where a = 1 − 40kλ h µh
l
. The linear term The pressure and velocity contours for two flows with
on the right-hand side, aQ, mainly denotes the pressure drop ReG = 12.8 and 127.6 are depicted in Figs. 10 and 11, respec-
required to overcome the viscous resistance. It is the contri- tively. In Fig. 10(a), it is seen that the pressure decreases
bution of the diffusion term of the Navier-Stokes equation. monotonically along the channel for the low ReG flow (lam-
The second term, bQ2 , is primarily the pressure difference inar). For the flow of ReG = 127.6 [Fig. 10(b)], however, the
needed to overcome the inertial effect, which is caused by lowest pressure takes place right after the throat of the channel.
the nonlinear convective term of the Navier-Stokes equa- This causes vortices, as indicated by the circular streaklines in
tion.15,38 The solid lines in Fig. 3 represent the predictions Figs. 6(d) and 6(j) [the vortices in Fig. 6(d) can be found in the
of Eq. (12). video in the supplementary material]. It is also seen that the

∗ S as a function of Re .
FIG. 8. Variation of Rin m G FIG. 9. Comparison between numerical simulations and experimental results.
112004-6 Tao et al. Phys. Fluids 30, 112004 (2018)

FIG. 10. Pressure changes. (a) Q = 4


ml/h and ReG = 12.8. (b) Q = 40 ml/h
and ReG = 127.6. In (a) and (b), the
top and middle panels show the pressure
contours and the bottom panel plots the
pressure variation along the centerline
of the channel.

largest pressure loss occurs around the throat for both flows, CONCLUSIONS
which presents a major contribution to the flow resistance,
We have investigated fluid flows in CDMCs. By consid-
especially for high ReG flows.
ering the geometry effect, a new dimensionless number Gm is
proposed and the product of Re and Gm, i.e., ReG , is shown
to be an appropriate dimensionless number for flow charac-
terization. Furthermore, a general law has been developed for
the estimation of flow resistance for transitional and turbulent
flows, which suggests a polynomial relationship between the
pressure drop and the flow rate. The dependence of the pres-
sure drop on the flow rate has also been confirmed by numerical
simulations.

SUPPLEMENTARY MATERIAL
See supplementary material for the following video: Flow
in channel IV as the flow rate is increased.

ACKNOWLEDGMENTS

FIG. 11. Velocity contours for the flows in Fig. 10. (a) Q = 4 ml/h and ReG This work was supported by the Research Grants Council
= 12.8. (b) Q = 40 ml/h and ReG = 127.6. of the Hong Kong Special Administrative Region under Grant
112004-7 Tao et al. Phys. Fluids 30, 112004 (2018)

No. 16210615. We thank the Nanosystem Fabrication Facility 19 T. Nishimura, Y. Bian, Y. Matsumoto, and K. Kunitsugu, “Fluid flow and
at the Hong Kong University of Science and Technology for mass transfer characteristics in a sinusoidal wavy-walled tube at moderate
Reynolds numbers for steady flow,” Heat Mass Transfer 39, 239 (2003).
the device fabrication. 20 A. M. Guzmán and C. H. Amon, “Dynamical flow characterization of tran-

sitional and chaotic regimes in converging–diverging channels,” J. Fluid


1 M. W. J. Prins, W. J. J. Welters, and J. W. Weekamp, “Fluid control in Mech. 321, 25 (1996).
multichannel structures by electrocapillary pressure,” Science 291, 277 21 D. F. Young, “Fluid mechanics of arterial stenosis,” J. Biomech. Eng. 101,
(2001). 157 (1979).
2 T. M. Squires and S. R. Quake, “Microfluidics: Fluid physics at the nanoliter 22 S. S. Varghese, S. H. Frankel, and P. F. Fischer, “Direct numerical simulation
scale,” Rev. Mod. Phys. 77, 997 (2005). of stenotic flows. Part 1. Steady flow,” J. Fluid Mech. 582, 253 (2007).
3 S. S. Hsieh and J. H. Liou, “DNA molecule dynamics in converging– 23 M. D. Griffith, T. Leweke, M. C. Thompson, and H. Kerry, “Effect of
diverging microchannels,” Biotechnol. Appl. Biochem. 52, 29 (2009). small asymmetries on axisymmetric stenotic flow,” J. Fluid Mech. 721, R1
4 J. Forrester and D. Young, “Flow through a converging–diverging
(2013).
tube and its implications in occlusive vascular disease–II: Theoretical 24 J. Samuelsson, O. Tammisola, and M. P. Juniper, “Breaking axi-symmetry in
and experimental results and their implications,” J. Biomech. 3, 307 stenotic flow lowers the critical transition Reynolds number,” Phys. Fluids
(1970). 27, 104103 (2015).
5 D. F. Young and F. Y. Tsai, “Flow characteristics in models of arterial 25 G. Leneweit and D. Auerbach, “Detachment phenomena in low Reynolds
stenoses—I. Steady flow,” J. Biomech. 6, 395 (1973). number flows through sinusoidally constricted tubes,” J. Fluid Mech. 387,
6 J. Vétel, A. Garon, D. Pelletier, and M. I. Farinas, “Asymmetry and transition
129 (1999).
to turbulence in a smooth axisymmetric constriction,” J. Fluid Mech. 607, 26 P. S. Mendes and E. M. Sparrow, “Periodically converging-diverging tubes
351 (2008). and their turbulent heat transfer, pressure drop, fluid flow, and enhancement
7 X. Xuan, B. Xu, and D. Li, “Accelerated particle electrophoretic motion
characteristics,” J. Heat Transfer 106, 55 (1984).
and separation in converging−diverging microchannels,” Anal. Chem. 77, 27 M. Akbari, D. Sinton, and M. Bahrami, “Viscous flow in variable cross-
4323 (2005). section microchannels of arbitrary shapes,” Int. J. Heat Mass Transfer 54,
8 X. Xuan and D. Li, “Particle motions in low-Reynolds number pressure-
3970 (2011).
driven flows through converging–diverging microchannels,” J. Micromech. 28 M. Akbari, D. Sinton, and M. Bahrami, “Laminar fully developed flow in
Microeng. 16, 62 (2005). periodically converging–diverging microtubes,” Heat Transfer Eng. 31, 628
9 S. S. Hsieh and Y. C. Huang, “Passive mixing in micro-channels with geo-
(2010).
metric variations through µPIV and µLIF measurements,” J. Micromech. 29 M. Akbari, D. Sinton, and M. Bahrami, “Analytical and experimen-
Microeng. 18, 065017 (2008). tal characterization of flow in slowly-varying cross-section microchan-
10 E. Lauga, A. D. Stroock, and H. A. Stone, “Three-dimensional
nels,” in ASME 2010 3rd Joint US-European Fluids Engineering Summer
flows in slowly varying planar geometries,” Phys. Fluids 16, 3051 Meeting collocated with 8th International Conference on Nanochannels,
(2004). Microchannels, and Minichannels (ASME, 2010), p. 2161.
11 J. S. H. Lee, Y. Hu, and D. Li, “Electrokinetic concentration gradient gener- 30 M. Akbari, D. Sinton, and M. Bahrami, “Flow in slowly varying microchan-
ation using a converging–diverging microchannel,” Anal. Chim. Acta 543, nels of rectangular cross-section,” in ASME 2009 Fluids Engineering
99 (2005). Division Summer Meeting (ASME, 2009), p. 1335.
12 S. Hardt, K. Drese, V. Hessel, and F. Schoänfeld, “Passive micromixers for 31 C. Duan and A. Majumdar, “Anomalous ion transport in 2-nm hydrophilic
applications in the microreactor and µTAS fields,” Microfluid. Nanofluid. nanochannels,” Nat. Nanotechnol. 5, 848 (2010).
1, 108 (2005). 32 L. Li, J. Mo, and Z. Li, “Flow and slip transition in nanochannels,” Phys.
13 T. J. Ober, S. J. Haward, C. J. Pipe, J. Soulages, and G. H. McKinley,
Rev. E 90, 033003 (2014).
“Microfluidic extensional rheometry using a hyperbolic contraction geom- 33 L. Li, J. Mo, and Z. Li, “Nanofluidic diode for simple fluids without moving
etry,” Rheol. Acta 52, 529 (2013). parts,” Phys. Rev. Lett. 115, 134503 (2015).
14 M. S. N. Oliveira, L. E. Rodd, G. H. McKinley, and M. A. Alves, 34 B. Ničeno and E. Nobile, “Numerical analysis of fluid flow and heat transfer
“Simulations of extensional flow in microrheometric devices,” Microfluid. in periodic wavy channels,” Int. J. Heat Mass Transfer 22, 156 (2001).
Nanofluid. 5, 809 (2008). 35 G. Wang and S. P. Vanka, “Convective heat transfer in periodic wavy
15 V. S. Duryodhan, S. G. Singh, and A. Agrawal, “Liquid flow through con-
passages,” Int. J. Heat Mass Transfer 38, 3219 (1995).
verging microchannels and a comparison with diverging microchannels,” 36 M. Akbari, D. Sinton, and M. Bahrami, “Pressure drop in rectangular
J. Micromech. Microeng. 24, 125002 (2014). microchannels as compared with theory based on arbitrary cross section,”
16 J. Q. Yong and C. J. Teo, “Mixing and heat transfer enhancement in
J. Fluids Eng. 131, 041202 (2009).
microchannels containing converging-diverging passages,” J. Heat Transfer 37 M. Bahrami, M. M. Yovanovich, and J. R. Culham, “Pressure drop of
136, 041704 (2014). fully-developed, laminar flow in microchannels of arbitrary cross-section,”
17 J. H. Forrester and D. F. Young, “Flow through a converging-diverging
J. Fluids Eng. 128, 1036 (2006).
tube and its implications in occlusive vascular disease—II: Theoretical and 38 Z. Zeng and R. Grigg, “A criterion for non-Darcy flow in porous media,”
experimental results and their implications,” J. Biomech. 3, 307 (1970). Transp. Porous Media 63, 57 (2006).
18 M. Faghri and Y. Asako, “Numerical determination of heat transfer and 39 C. Liu and Z. Li, “On the validity of the Navier-Stokes equations for

pressure drop characteristics for a converging–diverging flow channel,” nanoscale liquid flows: The role of channel size,” AIP Adv. 1, 032108
J. Heat Transfer 109, 606 (1987). (2011).

You might also like