Está en la página 1de 11

Sensors and Actuators B 139 (2009) 637647

Contents lists available at ScienceDirect

Sensors and Actuators B: Chemical


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

Numerical and experimental mixing studies in a MEMS-based


multilaminated/elongational ow micromixer
John T. Adeosun , Adeniyi Lawal
New Jersey Center for MicroChemical Systems, Department of Chemical Engineering and Materials Science, Stevens Institute of Technology,
Castle Point on Hudson, Hoboken, NJ 07030, USA

a r t i c l e

i n f o

Article history:
Received 2 February 2009
Received in revised form 10 March 2009
Accepted 16 March 2009
Available online 27 March 2009
Keywords:
Micromixer
Microfabrication
Passive mixing
Elongational ow
CFD
Residence time distribution

a b s t r a c t
Improvement of mixing quality in microchannel mixers or reactors has been recognized as a relevant
technical issue critical to the development and application of integrated microchemical processing systems. Silicon micro-electromechanical systems (MEMS) technology was successfully used to fabricate
a novel multichannel micromixer. This improved micromixer design basically used the mechanisms of
uid multilamination, elongational ow, and geometric focusing for mixing enhancement. The fabricated
triple-stack (PyrexTM /silicon/PyrexTM ) multilaminated/elongational ow micromixer (herein referred to
as MEFM-4) was evaluated for its mixing performance using residence time distribution (RTD) measure
in conjunction with UVvis absorption spectroscopy detection technique. Using a semi-empirical model
and the so-called convolutiondeconvolution theorem, a model description of the experimental RTD data
was obtained for the ow/mixing unit. This result was compared with numerical RTD predictions based
on computational uid dynamics (CFD) simulations. The simulation results are in good agreement with
the experimental data, especially in the low ow-rate range (Reynolds number <13 in this study). However, as expected, the accuracy of the CFD simulations is generally limited at higher ow rates (high Peclet
number) because of unavoidable numerical diffusion. This paper describes the efcient design, fabrication
and characterization of an effective microchannel mixer for microchemical systems applications.
2009 Elsevier B.V. All rights reserved.

1. Introduction
Mixing in microchannel mixers or reactors is one of the
key scientic issues that need to be fundamentally understood and technically addressed for microchemical systems to
take full advantage of their established and potential benets.
The microchannel-based process devices possess extremely high
surface-to-volume ratio (e.g. 40,000 m2 per 1 m3 in principle for
a simple rectangular channel with a cross-section of 100 m by
100 m), because of their small linear dimensions, a characteristic
from which derives most of their advantages over conventional-size
chemical process equipment. These advantages include high heat
and mass transfer rates (with shorter residence time) that enable
reactions to be performed under hitherto aggressive conditions
with higher yields and greater selectivity. Other key advantages
demonstrated or envisioned for certain applications of miniaturized chemical systems are: processing and hold-up safety for
extremely hazardous or toxic chemicals [1,2]; on-demand/on-site
synthesis of some critical chemicals such as H2 O2 [3]; better process control for effective processes/materials screening [4], easier

Corresponding author. Tel.: +1 201 216 5539; fax: +1 201 216 8306.
E-mail addresses: jadeosun@stevens.edu, jtadeosun@yahoo.com (J.T. Adeosun).
0925-4005/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.snb.2009.03.037

scale-up via numbering-up approach for exible production [5];


and the development of integrated chemical analytic platforms as
in micro total analysis systems (TAS) [6].
Indeed, the understanding and hence proper evaluation of the
ow and mixing behavior in microchemical processing systems
such as microchannel mixers and reactors is critical to their effective design and optimization. In combination with low processing
ow rate, the small transverse dimensions in microchannels imply
low Reynolds number (Re) laminar ow (Re typically between
0.1 and 100). As a consequence of low Re, mixing (without
enhancement) in microchannels would occur predominantly by
molecular diffusion which is very slow compared to convection.
Some extensive reviews [7,8] on passive and active microchannel
mixers have been published with interesting results. Hessel et al.
[9] carefully reviewed more than 30 types of passive micromixers (compared to 13 types of active micromixers) that used various
mechanisms or strategies for mixing enhancement. Such mixing
approaches include for instance: contacting of two substreams at a
T-junction [10]; geometric-focusing bi-laminated ow/mixing [11];
bifurcation [12] and interdigital [5,13] multilaminated ow/mixing;
split-and-recombine ow/mixing [14]; and chaotic mixing via
ridges-induced transverse ow [15]. Useful theoretical and/or
experimental studies were carried out on these micromixers (made
from various materials) with focus on their design and evaluation

638

J.T. Adeosun, A. Lawal / Sensors and Actuators B 139 (2009) 637647

for mixing enhancement. For passive or static mixers, the inherently high surface-to-volume ratio (with the associated high mass
transfer rate) in microscale mixer/reactor is generally made effective for mixing enhancement by engineering or manipulating the
geometrical channels and ow structures within this microscale
environment. Therefore, it is the technical know-how of imposing the needed geometric constraints with the associated mixing
mechanisms that distinguishes the mixing performance of one
micromixer from another.
The purpose of this research work is to design and characterize
the mixing behavior of a passive multilaminated/elongational ow
micromixer using both numerical and experimental approaches.
This work involves: using computational uid dynamics (CFD)
simulations as a vital design, optimization and characterization
tool; fabricating the mixing unit using silicon-MEMS technology;
and performing experiments to validate the numerical predictions
of the mixing characterization. In the present work, the proposed silicon-based multilaminated/elongational ow micromixer
combines at least three mixing mechanisms (discussed later in Section 2.1) for mixing enhancement. The numerical simulations and
experimental validations were performed using residence time distribution (RTD) as measure to characterize ow/mixing behavior
in the mixing-enhanced conguration. It is worth noting that the
choice and application of suitable characterization technique(s) for
the evaluation of mixing behavior in microchannels still pose some
challenges [8,16] that also need to be addressed.
Various mixing characterization measures/methods such as
ow visualization of test solutions (containing a chemical indicator)
and imaging [15,17], RTD [1820], and test chemical reactions (such
as VillermauxDushman competing parallel reactions) [5,21]
have been used by investigators to characterize theoretically and/or
experimentally the degree of mixing in microuidic devices. Some
of the characterization measures have been found particularly suitable for the qualitative analysis of mixing [17,21], but its application
to generate quantitative data for objective characterization of mixing can be limited by insufcient sensitivity of the test solutions and
the low local resolution of instrumentation [5]. The most desirable
characterization measures/methods should be those that not only
provide qualitative information for the comparisons of micromixers but also reliable quantitative data for direct characterization of
different micromixers.
The RTD measure chosen for this work, although a wellknown method [22] for characterizing ow and mixing behavior in
macromixers/reactors, is still a novel technique for the characterization of mixing in micromixers/reactors [1820,23]. By performing
the so-called tracer or stimulusresponse experiments, RTD data
that can be used for ow and mixing evaluation in continuous ow
systems can be obtained. The RTD data obtained from tracer experiments can be used directly or in combination with ow models
(depending on the ow system boundary conditions) to predict performance of non-ideal ow mixers/reactors. In the present study,
the RTD data were used in combination with a semi-empirical
RTD model to predict the ow/mixing behavior of the studied
micromixer. In the context of microchannels for low Re laminar
ow, the application of CFD tools in RTD studies becomes especially
useful when experimental RTD data can be obtained for comparison
and validation purposes.
The CFD tools have been effectively applied in our prior simulation studies [20,24] and by other researchers [2529] for
investigating ow and mixing behavior in microchannels for singleand multi-phase systems. These tools were used in our work
to obtain solutions to three-dimensional ow and mass transport equations, from which numerical RTD data were extracted.
A commercially available nite volume-based CFD software package, FLUENT , was used for our simulations, while all the other
numerical processing steps including numerical integration of

the RTD data were achieved using Mathematica software. The


present work is aimed at expanding the knowledge base for the
design, fabrication, and characterization of effective micromixers, the development of which will greatly benet the useful
applications of microchemical systems and other related novel
technologies.
2. Design and fabrication process
2.1. Design strategy for the proposed micromixer
In our earlier theoretical mixing study [20], three proposed
multilaminated/elongational ow micromixers (MEFMs) were
evaluated for their mixing performance. One of the MEFMs that
we referred to as MEFM-4 was selected for further study based
on the criteria of high mixing performance with minimum pressure drop. This best MEFM shown in Fig. 1 is herein referred
to as MEFM-4 because the mixing elements strategically placed
on the mixer channel oor are trapezoidal structures. Each of the
isosceles trapezoidal structures has a height of 857.8 m, two parallel sides of lengths 1000 and 1800 m, and an angle of 65
between the slant edge and the longer parallel side. For effective and rapid mixing enhancement, passive micromixers should
be designed such that the mixing mechanisms implemented in
them eventually lead to a reduction in diffusion path of uids in
the ow transverse direction and increase in uid-contact areas or
interfaces. With these goals in mind, the MEFM-4 was designed
based on the concept of uid multilamination and elongational
ow. The uid multilamination leads to the desired reduction in
uid diffusion path while the elongational ow causes an exponential increase in contact areas or interfaces for effective mass
transfer. As shown in Fig. 1, for intimate contact right at the onset
of feed inlets, uid A and uid B are introduced alternately into the
mixing channel. With this uids introduction approach, multilaminated streams of thin multilamellae of uids are formed with
the associated split-and-recombine ow downstream. Additionally, with the incorporation of the mixing elements, uid stretching
and local reorientation of uid interfaces results into the desired
elongational ow. In order to further enhance mixing in this conguration at a design level, the last ve rows of trapezoidal mixing
structures were ipped to form somewhat a mirror image of the six
upstream rows.
In order to attain high ow rates (Q) while maintaining a reasonable pressure drop, an appropriate ow distribution system
that will evenly distribute the ow into and out of the channels
of MEFM-4 (shown in Fig. 1) must be designed. Using computer-

Fig. 1. Multilaminated/elongational ow micromixer (MEFM-4).

J.T. Adeosun, A. Lawal / Sensors and Actuators B 139 (2009) 637647

639

2.2. Microfabrication procedure

Fig. 2. Model-based MEFM-4 with manifold design: (a) the frontside and (b) the
backside.

aided design (CAD) and CFD software packages, one uid-outlet


manifold and two uid-inlet manifolds one at the frontside and
the other at backside, were designed and optimized to facilitate the
uniform ow distribution of uids into the channels of MEFM-4 (see
Fig. 2 where L = 35.0 mm and W = 30.0 mm). The CFD tool used was
described in our earlier theoretical study [20]. Both uids/species
A and B are introduced from the inlet ports at the backside of the
device. The manifolds are designed such that uid A ows directly
to the front-side manifold (see Fig. 2a). Fluid B ows from the backside manifold (see Fig. 2b) and then through the four through-holes
(located along the line of conuence shown in Fig. 2a) to meet and
start mixing with uid A. The geometric focusing [30] nature of
this mixer design layout, i.e. two inlet streams splitting into eightinlet substreams (immediately after the line of conuence shown in
Fig. 2a) and recombining into four substreams downstream before
forming a single stream at the exit, also creates global reorientation of uid interfaces for effective mixing. This multichannel
conguration has a channel depth of 300 m and smallest owchannel dimension of 200 m resulting in a ow-domain volume
(V) of about 50 L.

The fabrication of our proposed MEFM-4 was successfully completed using the state-of-the-art equipment at Cornell NanoScale
Science and Technology Facility (CNF). The microuidic mixing unit
was made from silicon and PyrexTM using MEMS microfabrication technology and the associated microchannel fabrication and
packaging methods [31,32]. Specically applying silicon bulk micromachining techniques, which involve two main fabrication steps
namely photo-lithography (hereafter, lithography) and deep reactive ion etching (DRIE), the desired channels and structures were
created on both sides of a double-side polished silicon wafer (ptype 1 0 0 4-in. diameter, 800-m thick). The fabrication process
employed three steps of lithography and DRIE each in a predetermined sequence to ensure perfect back-to-frontside-alignment and
etching of the ow-through structures in MEFM-4 (see Fig. 2). As
expected, this alignment along with the associated etching process
was the most challenging requirement for the successful fabrication
of MEFM-4. The DRIE recipe enabled by inductively coupled plasma
(ICP) was utilized for the deep etching (with vertical walls) of the
mixing channels, manifold structures, and critical four through
holes in the MEFM-4. The mixing channels (on the frontside) and
the manifolds (on both the frontside and backside) were etched to
a depth of 300 m while the four ow-through holes (each with
a cross-sectional dimension of 200 m by 500 m) were etchedthrough from the backside of the 800-m thick wafer. More details
on the lithography and DRIE steps are described next.
For the required lithography steps, the CAD layouts of the
frontside and backside of MEFM-4 (with suitable alignment
marks) were rst transferred using an optical-pattern generator (GCA/MANN PG 3600F) and a mask-develop-etch equipment
(Hamatech HMP 900) to three 5-in.2 chrome-coated glass masks.
These three glass masks are herein referred to as M1 (comprising
the backside manifold pattern), M2 (comprising the pattern for the
four ow-through holes, inlet hole/opening to the backside manifold, inlet ow-through hole to the frontside manifold, and outlet
ow-through hole), and M3 (comprising the pattern for the channels and manifolds on the frontside). Fig. 3 shows the major steps
that are involved in the fabrication of MEFM-4. The schematic gures show basically the cross-section where the four through-holes
are located with the cross-section for the inlet and outlet holes on
MEFM-4 (see Fig. 2a). These steps are described herein fully. The
silicon wafer was initially coated (using thermal oxidation furnace)
with 2-m thick silicon dioxide, which would act as hard mask
needed later for the second DRIE step. In the rst lithographic step,
the pattern on the rst mask (M1) was transferred to the backside of
silicon/silicon-dioxide wafer (earlier spin-coated with a thick positive photoresist) by exposing the mask to UV light via a contact
aligner (EV 620). The oxide in the developed photoresist (PR) area
on the patterned wafer was then removed using a dielectric etcher
(Oxford 100). The PR on the wafer surface was stripped off using
two hot PR-stripper baths. The second mask (M2) was used in the
next lithographic step with the above lithography process repeated
on the backside of the wafer without removing the PR on the wafer
surface as in the last step above. The exposed silicon area (on the
wafer backside) was then etched to a depth of 200 m using DRIE
by inductively coupled plasma (Unaxis SLR-770). The etch rate was
approximately 2 m/min using this system. This DRIE step was for
etching the uidic structures (based on the pattern transferred from
M2) to a certain depth such that the remaining depth to be etched
would complete the etching requirement of the structural patterns
on both M2 and M1 for MEFM-4. The PR on the wafer surface (covering the pattern transferred from M1) was stripped off and the
exposed silicon area was etched 300 m deep with the 2-m thick
oxide hard mask. This deep silicon etching completes the processing
of the backside of the wafer.

640

J.T. Adeosun, A. Lawal / Sensors and Actuators B 139 (2009) 637647

etch-through of some of the backside structures, such as the critical four ow-through holes, one inlet ow-through hole, and one
outlet ow-through hole. The PR and silicon dioxide remnants on
the wafer were nally removed using two hot PR-stripper baths
and BOE solution, respectively.
After cleaning this wafer, the next step was to cover the structured wafer on both sides with 500-m thick PyrexTM glass wafers.
Fig. 4a and b shows the structured frontside and backside of one
of the four mixing units obtainable on a wafer. Using the Stationary Sonic-Mill Process (Stationary Model AP-1000), 800-m
diameter holes were drilled through one PyrexTM wafer for uidic
access to the structured silicon wafer. Two-step anodic bonding
method was used to cover the silicon wafer with PyrexTM wafers,
which facilitates optical and uidic access into the channels of the
micromixers. Using EV 501 wafer bonder with the EV 620 contact
aligner (for aligned bonding) at CNF, the structured PyrexTM was
anodically bonded to the bottom/backside of the silicon wafer. The
anodic bonding of the second PyrexTM wafer to the top/frontside
of the silicon wafer was carried out at Applied Microengineering Ltd. (AML). The last processing step that made the MEFM-4
complete for our mixing study was dicing the bonded triple-stack
(Pyrex/silicon/Pyrex) wafer into four individual micromixers of size
35.0 mm 30.0 mm each.
3. Numerical analysis
3.1. Mixing characterization measure

Fig. 3. Schematic gures showing the fabrication steps: (a) for the backside of the
Si wafer; and (b) for the frontside of the Si wafer and anodic bonding of the wafer
with Pyrex wafers.

The remaining silicon dioxide on the wafer was then removed


using a buffered oxide etch (BOE; 6:1 HF solution) reaction with
the oxide for about 30 min. Using thermal deposition system (GSI
PECVD), a 2-m thick oxide was deposited as hard mask on the
unprocessed frontside of the wafer while on the backside of the
wafer a 3-m thick oxide was deposited to act as etch stop for
the last deep etching process (from the frontside). Lithography was
then employed with the third mask (M3) to pattern the frontside
of MEFM-4 using 8-m thick PR. Using the dielectric etcher, the
oxide in the developed PR area on the patterned wafer was removed
in preparation for the last DRIE step. The exposed silicon area was
then etched to a depth of 300 m to obtain the frontside uidic
structures with the required depth of 300 m and consequently the

In choosing the mixing characterization techniques that are


applicable numerically and experimentally to microscale mixing
congurations, the following issues need to be considered: (i)
numerical diffusion, (ii) computational time, (iii) sensitivity of the
test solutions or reaction(s), (iv) the appropriate methods for the
accurate feeding/injection of the test uids, (v) the reliable detection of measurable signals in the outlet mixture or product(s)
stream, (vi) material of construction of the micromixer/detection
interface, and (vii) cost and relative ease of construction of the
experimental setup.
RTD is chosen for this work because it serves as a reasonable
indicator of the type and extent of mixing. In order to analyze and
characterize the mixing performance in real reactors, the concept
of RTD was rst extensively used by Danckwerts [33] and later discussed in details by some authors [3436]. The RTD can be obtained
by injecting a tracer instantaneously (a pulse input) or at a constant rate (a step input) at the inlet of a ow system, and then
measuring the tracer concentration, C(t), at the exit as a function
of time. A number of factors such as channeling, stagnation in dead
zones, axial dispersion, and imperfect mixing usually determine the
shape of the obtained response curves from the RTD experiments.
Danckwerts [33] dened a function, known as RTD function, which
describes quantitatively how much time different uid elements
have spent (i.e., distribution of the times spent) in a continuous
ow system. Mathematically, the RTD (or exit age-distribution)
function, E(t), can be dened such that for pulse input, E(t) is
related to the outlet tracer concentration, C(t), by the expression
given by:
E(t) =

C(t)

C(ti )

C(t)dt
0

(1)

C(ti )ti

i=0

where ti (= ti+1 ti ) is the time step for the measurements.


However, when the mixing performance of ow systems of different sizes or ow conditions is to be compared, a normalized RTD
function, E(), is used instead of E(t). Both functions are related

J.T. Adeosun, A. Lawal / Sensors and Actuators B 139 (2009) 637647

641

Fig. 4. Fabricated MEFM-4 (without PyrexTM covers): (a) the frontside and (b) the backside.

by Eqs. (2a)(2b), where tm is the mean residence time of the


RTD.
E() = tm E(t);

(2a)

t
tm

(t tm )2 E(t)t

(4)

t=0


tm

(5)

(2b)
3.2. Computational uid dynamics (CFD) simulation

RTD analysis [22] is generally applicable to ow system with


one-inlet stream (where tracer is injected). This limitation is
mitigated for our two-inlet, one-outlet ow system by injecting
the tracer into one of the two inlets (referred to as the maininlet) while feeding the second inlet with 1% of the main-inlet
water ow rate. This makes this ow system effectively a oneinlet, one-outlet system as desired without any ow disturbance.
The concentrationtime data obtained from a stimulusresponse
experiment can represent or be used to obtain the RTD provided
certain inlet and outlet boundary conditions earlier discussed [24]
are met. It is worth mentioning that for the comparison between
the CFD simulations and the laboratory experimental RTD data to
be valid, the CFD simulations should approximate closely the experimental methods of injection and measurement [37,38]. Once the
RTD function is obtained, statistical parameters such as the mean
residence time (tm ), variance ( 2 ) or square of the standard deviation, and coefcient of variation (CoV) can be obtained using Eqs.
(3), (4), and (5), respectively. The RTD of a microchannel mixer will
deviate from an ideal plug-ow mixer/reactor depending on the
hydrodynamics within the microchannel. In the context of the static
microchannel mixers designed to effect radial/transverse mixing, a
CoV of zero would imply complete plug-ow mixing while a nonzero CoV implies that there is axial dispersion or mixing caused by
non-uniform or laminar velocity prole and molecular diffusion. In
this case, the smaller the variance or the CoV, the narrower is the
RTD, the closer is the distribution to the mean residence time, and
the better the mixing quality.


tE(t)dt
tm =

(t tm )2 E(t)dt
=

 =

CoV =

where dimensionless time:


=


2

E(t)dt


E(t)dt = 1)

(since
0


t=0

( v) = 0

(6)

1
v
+ v v = p +  2 v

t

(7)

cA
+ (v. )cA = DAB 2 cA ;
t

tE(t)dt
=

CFD is generally recognized as a powerful tool for obtaining


the numerical solution to the equations of conservation of mass,
momentum, energy and chemical species describing the problem
of interest in a given ow geometry. Theoretically, the velocity and
concentration elds of a tracer, which can be obtained from the
solution of the transport phenomena equations [39], constitute all
the information that is needed to determine RTD in a continuous
ow system. For the theoretical analysis of the isothermal, laminar
mixing ow problem, the steady-state (or unsteady-state for RTD
analysis) ow and species mass transport of incompressible Newtonian uid(s) can be described by equations of conservation of
mass, momentum, and chemical species transport (for binary systems) given in Eqs. (68), respectively. In these equations, v is the
velocity vector, cA , A , and DAB are the concentration, mass fraction,
and binary diffusivity of the species A for system AB, respectively,
while p, , and  are the pressure, density, and kinematic viscosity
of the uid, respectively. In our simulation, uid/species A and B
represent tracer and water, respectively. By applying appropriate
boundary conditions to the geometrical conguration of interest,
the problem is specied completely and a unique solution can be
obtained.

tE(t)t

(3)

(8a)

where
cA = A

(8b)

Our numerical simulations were performed using a nite


volume-based commercial CFD code of FLUENT (Fluent 6.3.26
interfaced with Gambit 2.4.6) installed on a 64-bit master server
(Dual 3.0 GHz Intel Xenon with 8G of RAM and total hard-disk
size of about 1400 GB) within a Red Hat Enterprise Linux-cluster of
servers and workstations. Pro/Engineer , a powerful CAD software

642

J.T. Adeosun, A. Lawal / Sensors and Actuators B 139 (2009) 637647

package, was used for solid modeling of the micromixer conguration. GAMBIT preprocessor (CFD preprocessor from FLUENT )
was used for extracting the uid ow domains of the geometrical congurations and for meshing. As part of the measures to
obtain accurate solution there is need to have as many structured
elements (quadrilateral/hexahedral elements) as possible making
up the uid domain mesh and also aligning with the ow as perfectly as possible [40]. This was achieved by breaking the uid
domain into suitable sub-domains before meshing. In addition, a
mesh dependence study was carried out to determine the interval size of volume elements suitable for adequate mesh resolution
needed to obtain mesh-independent solution. Based on this study,
a mesh with 3.8 million nodes was found adequate (considering
computational expenses) and then exported from GAMBIT preprocessor to the FLUENT solver. The FLUENT segregated pressure-based
solver was used for the solution of three-dimensional uid ow
and species transport problem. The validity of the CFD code for
the simulation of RTD in microchannels has been examined by
comparing CFD simulation result in tubular microchannel with
the existing theoretical prediction [24]. In essence, the uid ow
equations and species transport equation were solved in a sequential manner using appropriate boundary conditions and numerical
algorithms.
The converged solution to steady state uid ow problem was
obtained rst. A uniform velocity prole was specied indirectly
(since density is constant) by setting the desired mass ow rate at
the inlet zone(s). The no-slip condition was specied as boundary
condition at the walls and gauge pressure of zero at the outlet zone
of the congurations. The properties specied for water (as a model
uid) were  = 998.2 kg/m3 and  = 1.005 106 m2 /s. In order to
obtain a stable, converged solution, certain parameters and convergence criteria such as under-relaxation factors and residuals
were set appropriately.
Using the converged solution of the steady state uid ow
equations, the tracer species equation was solved as an unsteady
simulation, whose solution was then used for RTD analysis. The
zero diffusive ux was specied as boundary condition at the walls
for the tracer species transport equation. Species mass fractions
of one (for time t = 0 at pulse injection of tracer into water for
rst time step) and switched to zero (at t > 0 after pulse injection
for the second and later time steps) were specied at the inlet
zone. The properties of water were specied for the tracer since
theoretically the tracer is taken as a water-like uid for our RTD
simulations. A mass diffusivity, DAB , of 1.5 109 m2 /s was specied for the tracer-water (AB) system. It should be noted that
discretization of the convective terms in momentum and species
transport equations (Eqs. (7) and (8a-b) above) usually introduces
an error generally referred to as numerical diffusion. This numerical error can be considerably minimized [41] by using CFD-quality
ne mesh and choosing appropriate discretization (interpolation)
schemes. Hence, Green-Gauss node based was chosen as the gradient option while the QUICK and second-order upwind schemes
in FLUENT were used as the higher order interpolation methods for
the convective terms in Eqs. (7) and (8a-b), respectively. The 2ndorder implicit unsteady formulation in FLUENT was also specied
in solving the species equation to give a more robust and accurate
solution for the unsteady state simulation. The tracer concentration
data (acquired at the outlet zone using the postprocessing component of FLUENT ) were then weighted by outlet surface area. This is
done to obtain through-the-wall (or spatial average concentration)
measurement [35], which closely represents the type of measurement obtainable at the outlet boundary of the ow/mixing system
in our tracer experiment. These weighted concentration data were
then exported into Mathematica for RTD analysis, in which tm ,
 2 , and CoV were obtained and used to evaluate the degree of
mixing.

4. Experimental analysis
4.1. Setup and procedure
The schematic of the setup for our tracer experiment is shown
in Fig. 5a. The experimental setup for mixing characterization in
micromixers using RTD measure was designed such that the timedependent tracer absorbance or concentration data can be obtained
at both the inlet and outlet of ow/mixing congurations of interest.
The silicon-fabricated MEFM-4 was mounted on a stainless-steel
block/stage custom-built for pressure-driven ow access into the
mixing unit (as shown in Fig. 5b). Suitable O-rings were used for
compression seal tting after aligning the three holes (two inlet
ports and one outlet port) at the backside of the mixing unit with
the corresponding holes in the block. Two syringe pumps (from KD
Scientic) with 10-mL syringes were employed to deliver constant
ow rates of deionized water via the two inlet ports into the mixing unit. A ow ratio of 100:1 was used (same as for simulation)
at the inlets in order to make the ow system effectively one-inlet,
one-outlet system with practically zero ow disturbance. A microvolume solution of uranine (a water-soluble tracer dye also known
as uorescein sodium), was then introduced as a pulse input into
the steady state ow of water. Through a 10-mL syringe on another
syringe pump, the tracer was introduced into the main inlet stream
using a computer-controlled four-port micro-volume sample injector (from Valco Instruments Co. Inc.). The detection system (from
Ocean Optics Inc.) used comprises miniature PC2000 PC plug-in
spectrometers with two congurations, a master and a slave, for
simultaneous detection and measurement of the tracer absorbance
at both the inlet and outlet sampling zones of the ow/mixing
device. The light source (tungsten halogen lamp), ow-through
sampling cells and the spectrophotometers were connected to
interact with one another using 400-m diameter optical bers
(see Fig. 5a).
The consistency and the linearity of the calibration curves from
calibration experiments for the two sampling regions show that
the tracer concentrations at which our RTD experiments were
carried out are within the linear BeerLambert response range
for the UVvis spectrophotometers. Therefore, the obtained tracer
absorbance, which is a quantity that is proportional to tracer
concentration (i.e. C(t) = kA(t), where k is the constant of proportionality) can be used directly after normalization for RTD analysis.
Based on the calibration experiment and initial tracer experimental runs, the optimum amount and concentration of tracer for RTD
experiments were determined.
At a time, say t = 0 s, a 1.0 L of 0.5 g/L (500 ppm) uranine
solution was injected for a very short time period of 0.145 s
into the owing deionized water while simultaneously initiating data acquisition using the spectrometer operating software.
The length of the tubing connecting the injection point to the
ow/mixing system is reduced to the minimum possible so as to
minimize the axial dispersion of the pulse before reaching the inlet
detection zone. The time-dependent absorbance data were then
acquired until the measurement time reached about ve times
the estimated values of (=V/Q), the theoretical average residence
time of uid in the microchannel. The absorbancetime curves
obtained from tracer experiment should be normalized by the
area under absorbancetime curve. The normalized absorbancetime data obtained were then used as concentrationtime data
for RTD analysis. Four replicates of experiments were performed
for each ow rate investigated with our experimental setup to
establish high repeatability of data. It should be noted that the
repeatability of data obtained via tracer experiment in microscale
ow systems depends largely on the ability to perform the
experiment under carefully controlled and optimal experimental
conditions.

J.T. Adeosun, A. Lawal / Sensors and Actuators B 139 (2009) 637647

643

Fig. 5. (a) Schematic diagram of the experimental setup. (b) MEFM-4 device mounted on a stainless steel block/stage.

4.2. Modeling of RTD


Needless to say, it is technically challenging to obtain the theoretical perfect pulse or step injection of tracer experimentally.
Therefore, the acquisition of the inlet tracer concentration data
is vital for the determination of RTD of microchannel ow systems while excluding the effects of the auxiliary components
from the measured cumulative output response. A mathematical
approach based on Convolution Integral theorem [35] is quite useful for extracting the model-based RTD (E(t
)). According to this
theorem, for a linear ow process in which a one-shot tracer injection (a typical imperfect pulse injection) was made, a relationship
exists between the time-dependent output tracer concentration
p
(i.e. Cout (t) obtained at time t), the E(t
), and the input tracer concentration (i.e. Cin (t t
) measured at time t
earlier than t). This is
given by the convolution integral:
p

t
Cin (t t
)E(t
)dt

=

Cout (t) =
0


t

Cin (t t
)E(t
)t

(9)

Applying the above numerical version of the convolution intem (t))


gral in Eq. (9), the measured output concentration data (i.e. Cout
can be mathematically tted with the predicted output concentrap
tion data, Cout (t), a convolution product of the input concentration
data and a suitable RTD model. The model description of RTD in a
ow system is useful for estimating parameters and hence the prediction of the ow/mixing behavior and/or the conversion for the
ow system.
Models that closely represent ow in the real system are
assumed in order to predict the performance of non-ideal reactors. Two RTD models: axial dispersion model (ADM) [19,35] and
a semi-empirical model (SEM) described by Ham and Platzer [42]

were used for RTD modeling in our prior work [24]. However, the
SEM was used in the present work since it was found to model
quite well the hydrodynamic behavior of the microchannel congurations studied better than the ADM. Boskovic and Loebbecke
[18] also used SEM and found it to be superior to ADM in their
work on modeling of RTD in microchannel mixers. The RTD function
based on the SEM [42], particularly suitable for the modeling of real
ow systems with certain asymmetric residence time distribution,
is shown in Eq. (10a-b):
E()
= E(t) E(t
)

=

MNtkN 
t N+1

t
tmax

N1 

tkN 

tN

t
tmax

N M1
(10a)

where
tk =

tmin tmax
tmax tmin

(10b)

In Eq. (10a) above, M and N are the model parameters with no


real physical signicance to ow or mixing behavior but the RTD
model obtainable is very useful in approximating RTDs with certain
asymmetry. In essence, the calculated mean and the variance of the
obtained RTDs can be used to obtain the CoV. The tmin and tmax are
the experimental minimum and the maximum residence times of
the tracer, which can be estimated from the output concentration
curves. In the convolutiondeconvolution technique, timedomain
curve tting via non-linear optimization was used to determine E(t
)
after estimating the parameters in the RTD model used. With the
assumed model, the objective of the optimization problem was to
obtain the values of parameters that minimize the sum of squares
of deviation () between the measured and the predicted outlet

644

J.T. Adeosun, A. Lawal / Sensors and Actuators B 139 (2009) 637647

Table 1
The results of the RTD analysis for CFD simulation and experiment at different ow conditions.
CFD simulation

Experiment

Flow rate (mL/min)

(s)

Re ()

Pef ()

P (Pa)

tm (s)

CoV (%)

Data points

tm (s)

CoV (%)

Data points

0.20
0.25
0.40
0.50

35.72
28.58
17.86
14.29

5.1
6.4
10.2
12.8

3420
4270
6840
8550

382
478
765
957

38.26
30.80
19.74
16.07

27.6
29.2
33.1
34.9

5857
4787
3192
2642

41.51
34.09
22.35
18.53

32.3
32.7
40.6
42.4

908
744
665
555

concentration curves. That is, from Eq. (9):

Based on calibration experiment and initial experimental runs,


a ow rate range of 0.200.50 mL/min was chosen for our stud-

ies. Hence, data were obtained at four different volumetric ow


rates (Q) 0.20, 0.25, 0.40, and 0.50 mL/min, for which the associated low Reynolds number range is 5.112.8. The Reynolds number
(Re = uLc /) and uid Peclet number (Pe = uLc /DAB ) are calculated
(shown in Table 1) based on the outlet cross-sectional area of
the microchannel mixer (1000 m by 300 m), the characteristic length/diameter (Lc ; based on the outlet ow cross-section),
and the average velocity (u) through this cross-sectional area.
Using the earlier-described experimental setup, time-dependent
concentration data were acquired (at the inlet and the outlet of
the mixing unit) automatically (starting from time t = 0) for the
pulse injection of uranine into the main-inlet water ow stream.
The implementation of the Mathematica 5.2 code for the required
convolutiondeconvolution of concentration and RTD model data led
to the determination of the predicted output concentration curve,
model parameters, and the associated RTD model. At ow rates of
0.20 and 0.40 mL/min, for instance, the predicted and the measured
output concentration curves based on SEM are shown in Fig. 6. As
expected, there is a better t of the experimental data at a lower
ow rate of 0.20 mL/min since from the model tting it has a lower

Fig. 6. The output concentration curves (predicted and the measured) based on SEM
for MEFM-4: (a) at 0.20 mL/min and (b) at 0.40 mL/min.

Fig. 7. (a) the extracted RTD curves (E(t


) vs. t) and; (b) the normalized RTDs (E vs.
) for MEFM-4 at ow rates of 0.20 mL/min and 0.40 mL/min.


=


m
Cout
(t)

t


2
Cin (t t
)E(t
)t

(11)

In order to solve this convolution and the resulting optimization problem for large number of data points obtained from our
experiments, a code was written and implemented using some programming functions available in Mathematica 5.2 software. After
performing code validation exercises, the RTD model parameters
M and N from SEM obtained based on the model tting was then
used to determine the RTD function (E(t
) vs. t) and normalized RTD
function (E vs. ) curves.
5. Results and discussion
5.1. Tracer experiment

J.T. Adeosun, A. Lawal / Sensors and Actuators B 139 (2009) 637647

645

Fig. 8. The comparison of the E vs.  curves for the CFD simulation and experiment at a ow rate of: (a) 0.20 mL/min; (b) 0.25 mL/min; (c) 0.40 mL/min; and (d) 0.50 mL/min.

sum of squares of deviation, = 0.00070, compared with = 0.00169


for Q = 0.40 mL/min. The extracted RTD curves (E(t
) vs. t) and the
normalized RTDs (E vs. ) obtained from the respective RTD data
are shown in Fig. 7a and b.

RTDs (E vs. ) obtained for MEFM-4 at the ow rates studied are
discussed further in the next sub-section.

5.2. CFD simulations

Based on through-the-wall measurements for the CFD simulations and experiment, the plots obtained for the normalized RTD
function (E ) as a function of dimensionless time () at the ow
rates of 0.20, 0.25, 0.40 and 0.50 mL/min for MEFM-4 are shown in
Fig. 8(a)(d), respectively. It can be seen from these gures that the
experimental normalized RTD curves are in good agreement with
the curves obtained from simulations.
Unlike a single continuous-ow stirred-tank reactor (CSTR) with
broad RTD [34], a series of identically sized (say, n) continuous-ow
stirred-tank reactors (n-CSTRs) are sometimes used to model nonideal tubular reactors so as to closely approach plug ow reactor
behavior (with narrower RTD) as n increases. In a similar manner, the proposed MEFM-4 is designed with mixing elements on
its channel oor and other mixing enhancement mechanisms to
improve ow/mixing in a way that the RTD becomes narrower. The
narrower the RTD, the closer it approaches ideal plug ow behavior.
It is not only necessary to qualitatively match the RTD curves but
to quantitatively compare the CoV values for the characterization
of the RTD.
The result of the RTD analysis for our simulation and experiment
are shown in Table 1. The moderate increase in CoV with increase
in ow rate implies a reasonable decrease in mixing performance
of MEFM-4. The same trend is obtained for both the experiment
and the simulation. Considering the operating laminar ow regime
(very low Re) in this study, the above results suggest that mixing
in MEFM-4 is being controlled by transverse (or radial) diffusion
such that mixing improves with residence time (and decrease in
Peclet number). Of course, there is optimum residence time for

Applying the CFD approach described in Section 3.2, the RTD


analysis was performed using FLUENT integrated postprocessor and
other relevant mathematical packages for the required statistical
treatment of the RTD data. The pressure drop result based on the
steady state simulation of ow in MEFM-4 is shown in Table 1.
Apart from higher throughput, lower pressure drop (P in a range
of 382957 Pa for the ow rates studied) is another advantage of
using multichannel mixing congurations compared with singlechannel congurations [20]. For the species transport simulation,
the spatial average concentration data (that reasonably represent
through-the-wall measurements) were obtained at the mixing unit
outlet instead of the ow-weighted average concentration data that
represent mixing-cup or closed-boundary measurements. This is as
a result of the fact that the through-the-wall measurements closely
mimic our experimental concentration data acquisition. Tens of
thousand time steps, depending on the ow rate or mean residence
time, were required to obtain convergence for each 3D simulation
run. The mean residence times (tm ) obtained at different ow rates
in Table 1 shows that the CFD predicts reasonably well the theoretical average residence times ( ). The degree of deviation: maximum
of 7% at lower ow rates (i.e. 0.20 and 0.25 mL/min) and about 12%
at higher ow rates (i.e. 0.40 and 0.50 mL/min), is expected. This
is because our mean residence time and RTD calculations via CFD
simulation were purposefully (as earlier mentioned) not based on
ow-weighted average concentration data, upon which theoretical
average residence time and RTD are derived [35]. The normalized

5.3. Comparison of experimental data with numerical predictions

646

J.T. Adeosun, A. Lawal / Sensors and Actuators B 139 (2009) 637647

good mixing. The real value of using CoV as a measure of mixing can be seen when different congurations are compared for
their mixing performance. In fact, it has been shown in our earlier
numerical work [20] that MEFM-4 exhibits remarkably better mixing performance than T-junction and other micromixers studied
(at same mean residence times) since its RTDs are narrower with
smaller CoV values. Caution need be exercised when interpreting
RTD data because obtaining narrow RTD does not necessarily imply
good mixing in every mixing/reaction system. However, if good
mixing is carefully designed into and optimized for a microchannel mixer or reactor, narrow RTD should be expected. The above
results show that the RTDs from CFD simulation can predict reliably
well the non-ideal ow behavior in microchannel mixers/reactors,
and can therefore be used in estimating conversion in microreactor
systems.
6. Conclusions
The mixing enhancement at low Re ow regime in a multilaminated/elongational ow micromixer (MEFM-4) was investigated
experimentally as well as numerically. The work involves the
design, numerical simulation, fabrication, and mixing characterization of this proposed multichannel micromixer. This study shows
that the design of an ideal passive micromixer or reactor should
be based on the concept of uid multilamination and elongational
ow such that high mixing performance and high throughput can
be combined with minimized pressure drop.
Tracer experiment utilizing UVvis absorption spectroscopy
detection technique was used to obtain the required concentration
data for RTD analysis. With the aim of obtaining model description of RTD in the MEFM-4 studied, a semi-empirical model (SEM)
was applied to the experimental tracer data. By performing numerical experiments using CFD tools, the laboratory acquisition of
RTD data in MEFM-4 was suitably mimicked. The obtained data
were used to determine the characteristic moments of RTD such as
mean residence time and variance (or CoV). These measures were
then used to indirectly characterize mixing behavior in MEFM-4.
The MEFM-4 shows a good mixing performance considering its
narrow RTD with the low values of CoV obtained at the operating low Re ow regime. Results of the comparison between
the normalized experimental RTDs and those from CFD simulation show good agreement at various ow rates (5 < Re < 13 in this
study). This paper also demonstrates the viability of the rapidly
developing silicon-based microfabrication technology for creating
effective microuidic mixer/reactor that can be integrated with
sensing and/or actuating functions for suitable applications in
(bio)chemical synthesis and analysis, micro total analysis systems,
and other microsystems.
Acknowledgements
The authors would like to thank The Department of EnergyIndustrial Technologies Program (DOE-ITP) and American Chemical
Society Petroleum Research Fund (ACS PRF) for the grants provided
in support of this research project. The silicon and glass microfabrication was performed in part at the Cornell NanoScale Facility
(CNF), a member of the National Nanotechnology Infrastructure
Network, which is supported by the National Science Foundation
(Grant ECS-0335765). We also gratefully acknowledge the following
individuals for their technical assistance and helpful discussions:
Meredith Metzler and Mike Skvarla (CNF), Prof. R. Besser (New Jersey Center for MicroChemical Systems NJCMCS), Dr. Dongying
Qian (formerly with NJCMCS; now with CBI Co., NJ), Nick Aitken
(Applied Microengineering Ltd., UK), and Dusan Boskovic (Fraunhofer ICT, Germany).

References
[1] S.J. Haswell, R.J. Middleton, B. OSullivan, V. Skelton, P. Watts, P. Styring, The
application of micro reactors to synthetic chemistry, Chem. Commun. 5 (2001)
391398.
[2] J.R. Burns, C. Ramshaw, Development of a microreactor for chemical production,
Chem. Eng. Res. Des. 77 (1999) 206211.
[3] Y. Voloshin, R. Halder, A. Lawal, Kinetics of hydrogen peroxide synthesis by
direct combination of H2 and O2 in a microreactor, Catal. Today 125 (2007)
4047.
[4] R. Srinivasan, I.M. Hsing, P.E. Berger, K.F. Jensen, S.L. Firebaugh, M.A. Schmidt,
M.P. Harold, J.J. Lerou, J.F. Ryley, Micromachined reactors for catalytic partial
oxidation reactions, AIChE J. 43 (1997) 30593069.
[5] W. Ehrfeld, K. Golbig, V. Hessel, H. Lowe, T. Richter, Characterization of mixing
in micromixers by a test reaction: Single mixing units and mixer arrays, Ind.
Eng. Chem. Res. 38 (1999) 10751082.
[6] J. West, M. Becker, S. Tombrink, A. Manz, Micro total analysis systems: latest
achievements, Anal. Chem. 80 (2008) 44034419.
[7] V. Hessel, H. Lowe, F. Schonfeld, Micromixersa review on passive and active
mixing principles, Chem. Eng. Sci. 60 (2005) 24792501.
[8] N.T. Nguyen, Z. Wu, Micromixersa review, J. Micromech. Microeng. 15 (2005)
R1R16.
[9] V. Hessel, L. Holger, A. Muller, G. Kolb, Chemical Micro Process Engineering: Processing and Plants, Ist ed., Wiley-VCH, Weinheim, 2005, pp. 1
280.
[10] D. Gobby, P. Angeli, A. Gavriilidis, Mixing characteristics of T-type microuidic
mixers, J. Micromech. Microeng. 11 (2001) 126132.
[11] T.T. Veenstra, T.S.J. Lammerink, M.C. Elwenspoek, A. Van Den Berg, Characterization method for a new diffusion mixer applicable in micro
ow injection analysis systems, J. Micromech. Microeng. 9 (1999) 199
202.
[12] F.G. Bessoth, A.J. DeMello, A. Manz, Microstructure for efcient continuous ow
mixing, Anal. Commun. 36 (1999) 213215.
[13] V. Hessel, S. Hardt, H. Lowe, F. Schonfeld, Laminar mixing in different interdigital micromixers: I. Experimental characterization, AIChE J. 49 (2003)
566577.
[14] N. Schwesinger, T. Frank, H. Wurmus, A modular microuid system with an
integrated micromixer, J. Micromech. Microeng. 6 (1996) 99102.
[15] A.D. Stroock, S.K.W. Dertinger, A. Ajdari, I. Mezic, H.A. Stone, G.M. Whitesides,
Chaotic mixer for microchannels, Science 295 (2002) 647651.
[16] Z. Yang, S. Matsumoto, H. Goto, M. Matsumoto, R. Maeda, Ultrasonic micromixer
for microuidic systems, Sens. Actuators A 93 (2001) 266272.
[17] T.N.T. Nguyen, M.-C. Kim, J.-S. Park, N.E. Lee, An effective passive microuidic mixer utilizing chaotic advection, Sens. Actuators B 132 (2008) 172
181.
[18] D. Boskovic, S. Loebbecke, Modelling of the residence time distribution in
micromixers, Chem. Eng. J. 135 (2008) S138S146.
[19] F. Trachsel, A. Gunther, S. Khan, K.F. Jensen, Measurement of residence
time distribution in microuidic systems, Chem. Eng. Sci. 60 (2005) 5729
5737.
[20] J.T. Adeosun, A. Lawal, Mass transfer enhancement in microchannel reactors by
reorientation of uid interfaces and stretching, Sens. Actuators B 110 (2005)
101111.
[21] S. Panic, S. Loebbecke, T. Tuercke, J. Antes, D. Boskovic, Experimental approaches
to a better understanding of mixing performance of microuidic devices, Chem.
Eng. J. 101 (2004) 409419.
[22] E.L. Paul, V.A. Atiemo-Obeng, S.M. Kresta, in: E.L. Paul, V.A. Atiemo-Obeng, S.M.
Kresta (Eds.), Handbook of Industrial MixingScience and Practice, John Wiley
& Sons, Hoboken, NJ, 2004, pp. xxxiiilxi.
[23] S. Lohse, B.T. Kohnen, D. Janasek, P.S. Dittrich, J. Franzke, D.W. Agar, A novel
method for determining residence time distribution in intricately structured
microreactors, Lab Chip Miniatur. Chem. Biol. 8 (2008) 431438.
[24] J.T. Adeosun, A. Lawal, Numerical and experimental studies of mixing characteristics in a T-junction microchannel using residence-time distribution, Chem.
Eng. Sci. 64 (2009) 24222432.
[25] J. Aubin, D.F. Fletcher, C. Xuereb, Design of micromixers using CFD modelling,
Chem. Eng. Sci. 60 (2005) 25032516.
[26] T. Glatzel, C. Litterst, C. Cupelli, T. Lindemann, C. Moosmann, R. Niekrawietz,
W. Streule, R. Zengerle, P. Koltay, Computational uid dynamics (CFD) software tools for microuidic applicationsa case study, Comput. Fluids 37 (2008)
218235.
[27] F. Schonfeld, S. Hardt, Simulation of helical ows in microchannels, AIChE J. 50
(2004) 771778.
[28] A.K. Heibel, P.J.M. Lebens, J.W. Middelhoff, F. Kapteijn, J. Moulijn, Liquid residence time distribution in the lm ow monolith reactor, AIChE J. 51 (2005)
122133.
[29] D. Qian, A. Lawal, Numerical study on gas and liquid slugs for Taylor ow in a
T-junction microchannel, Chem. Eng. Sci. 61 (2006) 76097625.
[30] S. Hardt, F. Schonfeld, Laminar mixing in different interdigital micromixers: II.
Numerical simulations, AIChE J. 49 (2003) 578584.
[31] M. Madou, Fundamentals of Microfabrication, 2nd ed., CRC Press, Boca Raton,
FL, 2002.
[32] E. Verpoorte, N.F. De Rooij, Microuidics meets MEMS, Proc. IEEE 91 (2003)
930953.
[33] P.V. Danckwerts, Continuous ow systems: distribution of residence times,
Chem. Eng. Sci. 2 (1953) 113.

J.T. Adeosun, A. Lawal / Sensors and Actuators B 139 (2009) 637647


[34] H.S. Fogler, Elements of Chemical Reaction Engineering, 3rd ed., Prentice Hall
PTR, Upper Saddle River, NJ, 1999, pp. 809918.
[35] O. Levenspiel, Chemical Reaction Engineering, 3rd ed., John Wiley & Sons, New
York, 1999, pp. 257349.
[36] E.B. Nauman, Residence Time Distributions, Wiley, New York, 1983, pp. 3
52 (Chapter 1).
[37] E.B. Nauman, Residence time theory, Ind. Eng. Chem. Res. 47 (2008) 3752
3766.
[38] O. Levenspiel, J.C.R. Turner, The interpretation of residence-time experiments,
Chem. Eng. Sci. 25 (1970) 16051609.
[39] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, 2nd ed., John Wiley
& Sons, New York, 2001, pp. 846848.
[40] FLUENT 6.3 Documentation, Fluent Inc., Lebanon, New Hampshire, 2006.
[41] E.M. Marshall, A. Bakker, in: E.L. Paul, V.A. Atiemo-Obeng, S.M. Kresta (Eds.),
Computational Fluid Mixing, John Wiley & Sons, New Jersey, 2004, pp. 257338
(Chapter 5).
[42] J.H. Ham, B. Platzer, Semi-empirical equations for residence time distributions
in disperse systemsPart 1: continuous phase, Chem. Eng. Technol. 27 (2004)
11721178.

647

Biographies
John T. Adeosun obtained his B.Sc. in Chemical Engineering (1998) from Obafemi
Awolowo University, Nigeria and his M.Eng. in Chemical Engineering (2004) from
Stevens Institute of Technology, Hoboken, New Jersey, where he would be obtaining his Ph.D. in Chemical Engineering (2009). His research interest is in the design,
fabrication, and characterization of microscale mixers/reactors using CFD, and
experimental study of the transport processes and reaction kinetics in these microchemical systems.
Adeniyi Lawal obtained his S.M. (1982) and Ph.D. (1985), both in Chemical Engineering, from Massachusetts Institute of Technology (MIT), Cambridge, and McGill
University, Canada, respectively. He is a Professor of Chemical Engineering at the
Department of Chemical Engineering and Materials Science, Stevens Institute of
Technology, Hoboken, New Jersey. His research interests are in mathematical modeling of transport processes in complex macro- and micro-geometries, and design and
demonstration of microreactor systems for on-demand, on-site chemical synthesis
and biofuel production.

También podría gustarte