You are on page 1of 13

Chemical Engineering Journal 223 (2013) 454–466

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

Controlling lateral nanomixing and velocity profile of dilute ferrofluid


capillary flows in uniform stationary, oscillating and rotating magnetic
fields
Pouya Hajiani, Faïçal Larachi ⇑
Chemical Engineering Department, Laval University, Québec, QC, Canada G1V 0A6

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Nanoparticle maneuvered in laminar This study utilizes magnetically excited rigid-dipole magnetic nanoparticles, seeded in nonmagnetic liq-
capillary flow using magnetic field. uids at very low volume concentration and subjecting the solution to different types of magnetic fields (a:
 Spinning nanoparticle in rotating field RMF, b: OMF, and c: SMF) to investigate the effects on transport properties of liquid in capillary flows.
transfers momentum to liquid.
 Oscillating field does not promote
nanoparticle momentum transfer to
liquid.
 Static field locks nanoparticle to resist
against momentum transfer to liquid.

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

Article history: The influence of magnetic-field dependent viscosity (rotational viscosity) on molecular transport of spe-
Received 29 January 2013 cies in dilute ferrofluids has been studied. For this purpose, a Taylor dispersion test in a capillary tube has
Received in revised form 25 February 2013 been performed while suspended magnetic nanoparticles (MNPs) are subjected to both magnetic field
Accepted 28 February 2013
and low Re shear flow field. Axial dispersion has been quantified from residence time distributions (RTDs)
Available online 14 March 2013
and tracer injection tests conducted in three distinct situations where the capillary is subjected to (a) uni-
form transverse rotating magnetic field (TRMF), (b) uniform transverse oscillating magnetic field (TOMF),
Keywords:
and (c) uniform axial static magnetic field (ASMF). The various types of magnetic fields have been gen-
Magnetic nanoparticle
Magnetic field (uniform stationary,
erated in a specially designed stator energized by three phase, AC and DC currents. Results obtained from
oscillating, rotating) the three cases are reported in terms of axial dispersion coefficients. For TRMF, an increase in lateral mix-
Taylor dispersion ing is observed whereas no significant effect is detected for TOMF. In ASMF, the lateral mixing mechanism
RTD test is retarded by magnetically locked MNPs. Both effects under TRMF and ASMF reach a plateau as MNP con-
Mixing centration in the liquid is increased. These findings highlight the effect of rotational viscosity on diffusion
Laminar velocity profile of other species hosted in dilute ferrofluids and point to attractive applications to engineering fields
where transport phenomena are central. Analysis of RTD breakthrough times enabled laminar velocity
profile in capillary flow to be reconstructed. It suggests that (magnetic field-free) parabolic velocity pro-
files evolve towards flattened and protruded shapes, respectively, in TRMF and ASMF. These results con-
firm that magnetically-excited MNPs may be considered as a potentially appealing tool to mediate
molecular transport phenomena at the nanoscale such as in nano/microfluidic systems.
Ó 2013 Elsevier B.V. All rights reserved.

⇑ Corresponding author.
E-mail address: faical.larachi@gch.ulaval.ca (F. Larachi).

1385-8947/$ - see front matter Ó 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cej.2013.02.129
P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466 455

Nomenclature

d capillary tube diameter, m u local linear velocity in capillary, m/s


dp volume median particle diameter, m U average linear velocity, m/s
dp-p average particle–particle distance, m
D molecular diffusivity, m2/s Greek
Deff effective diffusivity in presence of magnetically excited g0 dynamic viscosity (shear viscosity), Pa s
MNPs, m2/s h dimensionless time, t/s
Vh MNP hydrodynamic volume, m3 m kinematic viscosity, m2/s
H0 external magnetic field vector, A/m sB Brownian relaxation time constant, s
H magnetic field intensity, A/m sOMF oscillating magnetic field period = 1/XOMF, s
I current intensity, A sRMF rotating magnetic field period = 1/XRMF, s
Imax peak current intensity, A / MNP core volume fraction
kB Boltzmann’s constant, J/K v0 initial susceptibility
K axial dispersion coefficient, m2/s W surface current distribution, A/m2
L distance between two electrodes, m x MNP spin velocity per unit volume, rad/s
m magnetic dipole moment of single domain nanoparti- X frequency, Hz
cles, Am2
Ms saturation magnetization, A/m Subscripts
np the number of pole pairs in the stator windings cur electrical current
n power low index in non-Newtonian power-law radial MNP in presence of MNPs
velocity profile x directional component
Pe Péclet number, Ud/D z directional component
r radial position in cylindrical coordinate, m
R capillary tube radius, m Acronyms
Rp diffusing particle radius in Stocks–Einstein law for diffu- MNP magnetic nanoparticle
sion in solution OMF oscillating magnetic field
Re Reynolds’ number, Ud/m SMF static (DC) magnetic field
RTD residence time distribution ID inner diameter
t time, s rms root mean square
t capillary mean residence time, s RTD residence time distribution
tmin capillary minimum residence time, s VFD variable frequency derive
T absolute temperature, K

1. Introduction Rotational viscosity under OMF was investigated experimen-


tally by Bacri et al. [7] after it was conceived theoretically by Shli-
Stable colloidal suspensions of single-domain magnetic nano- omis and Morozov [8]. Under OMF, rdMNPs try to chase the
particles (MNPs) in aqueous or organic nonmagnetic liquid carri- magnetic field via synchronized rotation due to periodic alterna-
ers, commonly known as ferrofluids [1], exhibit fascinating tion of the magnetic field. Subjecting a ferrofluid laminar Poiseuille
rheological properties under the effect of external magnetic fields flow to low-frequency coaxial OMF (oscillating at frequency XOMF)
[2]. A magnetic field exerts a magnetic torque on the magnetic mo- such that XOMFsB < 1(sB = 3g0Vh/kBT is Brownian relaxation time)
ment of MNPs suspended in the liquid in an attempt to orient them partially impedes the azimuthal spin of MNPs under fluid vorticity
toward its own dominant direction [1]. For rigid dipole MNPs and results in a positive rotational viscosity manifesting as pres-
(rdMNPs), featuring a magnetic moment locked to the particle so- sure drop augmentation [9]. Conversely, under high-frequency
lid crystal structure, the magnetic torque is felt bodily thereby OMF such that XOMFsB > 1 the magnetic field amplifies azimuthal
transferring concomitantly momentum to the adjacent liquid MNP spin. Accordingly, electromagnetic energy is transferred to
phase [1,2]. This magnetic body torque is always opposed by the fluid flow as kinetic energy, thereby enhancing flow nearby
Brownian collisions from solvent molecules and by flow-field fric- MNPs and lessening pressure drop. This effect may be viewed as
tional torque when the liquid is out of rest [1,2]. In a ferrofluid if the apparent ferrofluid viscosity has decreased tantamount to
shear flow, it is known that an external magnetic field forces negative rotational viscosity [7,9,10]. Gauzeau et al. [11]
rdMNPs to spin asynchronously relative to the local vorticity of studied rotational viscosity in concentrated ferrofluids with large
the fluid, resulting in a directional modification of viscosity. This sB sB  0.01 s) under OMF and determined that rotational viscosity
behavior, known as rotational viscosity, was first observed by depends also on flow vorticity.
McTague [3] and then by Rosensweig et al. [4]. Rotational viscosity under RMF was first observed experimen-
Since then, rotational viscosity has been studied vastly under tally by Moskowitz and Rosensweig [12] in the form of ferrofluid
three types of magnetic fields, i.e., static (or DC) magnetic field entrainment by the field originating from the average rotation of
(SMF), oscillating magnetic field (OMF) and rotating magnetic field MNPs relative to their embedding matricial fluid. The velocity pro-
(RMF) [1,2,5]. SMF rotational viscosity arises when the external file of this macroscopic motion in RMF was studied experimentally
magnetic field is not parallel to local fluid vorticity. While the fluid and theoretically by Chaves et al. [13,14]. The ferrofluid rheological
viscid nature forces rdMNPs to rotate in shear flows, the normal- properties under RMF have been obtained from measurement of
to-vorticity component of magnetic field vector hinders the dipoles torques of a submerged spindle. Based on this method, Rosental
free rotation, thus leading to a greater rate of hydrodynamic energy et al. [15] and later, Rinaldi et al. [16] detected negative and posi-
dissipation around the MNPs [1]. This results in an inflated appar- tive rotational viscosity with co-rotating and counter-rotating
ent viscosity as observed by McTague [3] and Rosensweig et al. [4] magnetic fields with respect to spindle resistive torques. Interest-
and explained theoretically by Shliomis [6]. ingly, under RMF, no minimum field frequency is required to
456 P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466

achieve negative rotational viscosity in contrast with Bacri et al. [7]


a b
experiments reported under OMF.
Except research works on macroscopic momentum transfer in
liquids (concealed in apparent viscosity alterations) [13,17], other
ferrofluid transport properties under constant or time-varying
magnetic fields have thus far not received sufficient attention. Chil-
ton–Colburn analogy [18] states that heat, mass and momentum
transfer phenomena are subtended by the same basic mechanisms.
Inspired by such analogy, together with rotational viscosity effects
showcasing positive, zero and negative apparent viscosity in liquid,
the quest of this study is to investigate how magnetically-excited
MNPs, seeded in liquids, could manipulate other molecular trans-
port properties, in particular liquid mass transfer. Notably, dilute
ferrofluids (magnetic core volume fraction / below 0.01) [11] are Phase 1
Phase 3
of particular interest in this study, as they may lead to potential Phase 2
applications in chemical reaction engineering where MNPs are al- c
ready used as support-shuttles for recovery of costly catalyst com-
plexes [19–24].
Aiming at the same objective, Suresh and Bhalerao [25] at-
tempted to enhance mass transfer at the gas–liquid interface of
two-phase flow systems. Ferrofluids were added to alkaline aque-
ous solutions in wetted-wall falling-film contactor and bubble col-
umn to sense the effect of MNP addition on CO2 absorption under
OMF for a diffusion-limited reaction. A 50-Hz longitudinal OMF
(LOMF) generated by two coils (mounted one on top of the other)
led to 50% enhancement of mass transfer. More recently, Komati
and Suresh [26] used a 50% v/v blend of ferrofluid (/ = 0.004)
and MDEA solution to study CO2 absorption in a falling-film cont- Fig. 1. Taylor dispersion in capillary tube exposed to magnetic field. (a) Schematic
actor exposed to LOMF. MNPs excited by LOMF from 100 to drawing depicts dispersion of a tracer blob in Poiseuille flow with parabolic laminar
1000 Gauß and from 500 Hz to 100 kHz led to almost doubled vol- velocity profile. (b) Schematic of the experimental setup including two-pole three-
phase magnet and glass-made capillary tube at the center. (c) Upfront view of
umetric mass transfer coefficient.
magnet with a capillary set vertically and coaxially with magnet bore, a uniform
In this contribution, we perform Taylor dispersion tests in a horizontal magnetic field imposed across capillary tube hosting a flow of MNP-laden
capillary tube and determine the axial dispersion coefficient (K) suspension.
as a means to investigate liquid-phase mass transfer in the
presence of magnetically excited MNPs interacting with shear flow.
Taylor dispersion is a phenomenon originating from combination The resistance and inductance of the stator windings were
between axial convective displacement of a tracer slab under lam- measured as 14.6 X per coil (single winding) and 150 mH per coil
inar Poiseuille flow and its diffusive migration in radial direction at 200 Hz.
[27]. Hence in a given experimental condition, any phenomenon To generate a RMF transverse to capillary flow such as in Fig. 2a,
that alters molecular diffusion could be detected indirectly through coils are fed by three balanced AC currents, 120° out of phase
its effect on K [27]. Residence time distribution (RTD) measure- from a variable frequency drive (ABB, ACS150, 2.2 kW), providing
ments are thus performed using a Taylor dispersion capillary cell z-directed surface current distribution, Kz, given by
from which axial dispersion coefficients are obtained. MNPs in 
Wz ðh; tÞ ¼ Re WejðXcur tnp hÞ ð1Þ
our Poiseuille experiments were excited successively by RMF,
OMF and SMF to probe their influence on the lateral molecular where Xcur is the electrical current frequency and np is a positive
transport mechanism in the capillary flows. Also, we put forward integer parameter representing the number of pole pairs in stator
an approach to correlate (and infer from impulse RTD) the shape (here, np = 3). In Fig. 2b, two adjacent coils are energized with an
of laminar velocity profile under magnetic fields from the mini- AC current from an AC variable frequency drive (Invertek Drives,
mum residence (or breakthrough) time of the capillary tube. Optidrive E2) to generate an OMF of Hx = H cos(xcurt) transverse
to capillary flow. Fig. 2c illustrates two adjacent coils connected
to a DC current from a DC source (Agilent Tech, N8739A) to provide
2. Experimental section an SMF parallel to capillary flow. The magnetic field strength and
frequency is adjusted directly by power supplies. The temperature
2.1. Magnet of the magnet solid part is controlled by a water cooling jacket that
encompasses the outer shell of the stator and filled with a coolant
A tubular two-pole three-phase magnet was designed and built circulated in and out from a constant-temperature thermostated
in collaboration with MotionTech LLC and Windings Inc. with bore bath (Lauda, Model RKT20).
dimensions as 55 mm height and 45 mm inner diameter. The
magnet assembly consists of three identical coil pairs, spatially 2.2. Taylor dispersion in capillary
shifted from each other by 120° in azimuthal direction as depicted
in Fig. 1a and b. Each coil can be energized separately or coupled to A Taylor dispersion test as first elaborated by Taylor [28,29]
other coils in various configurations. Therefore, the magnet uses a thin band of solute injected in a capillary tube in which a
can be used to generate different magnetic field types including laminar non-zero velocity field u(r) = 2U(1  (r/R)2) is dominant.
RMF, OMF and SMF (Fig. 2) with moderate intensity at the At time equal zero, the tracer band starts to be stretched by the
center axis (up to 50 mT). In absence of any magnetic object, each convective flow to evolve into a paraboloidal shell. The liquid
ampere rms generates a RMF of nearly 186 Gauß at bore center. velocity on the capillary wall is zero and hence, tracer molecules
P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466 457

Phase 1
Phase 3
a Phase 2
I/Imax
1.0
Phase 1
0.5 Phase 2
Phase 3

0.0
0 1/5 2/5 3/5 4/5 1

-0.5

-1.0
θ = t/τRMF

AC Open AC

b
1.0
I/Imax
0.5

0.0
0 1/5 2/5 3/5 4/5 1

-0.5

-1.0

θ = t/τ OMF

DC Open DC

c
1.0
I/Imax
0.5

0.0
0 1 2 3 4 5 6
-0.5

-1.0

θ = t/TDC

Fig. 2. Taylor dispersion capillary tube submitted to three magnetic field scenarios. Two-pole three-phase magnet generates (a) uniform rotating magnetic field (RMF) when
energized by a three-phase power supply, (b) oscillating magnetic field (OMF) when energized by a single-phase power supply, and (c) uniform static magnetic field (SMF)
when energized by a DC current. Perpendicular to (y, x) plane is z-direction.

close to centerline outpace those closer or adjacent to wall 2.3. Impulse RTD test
(Fig. 1a). Consequently, a radial concentration gradient forms over
the expanded stripe and tracer molecules diffuse radially to fade The axial dispersion coefficient in a vessel can be estimated
out the paraboloidal shape of the stretched band. Molecular diffu- through impulse RTD tests which, due to their simplicity and effec-
sion frees tracer molecules from being trapped in the wall vicinity tiveness, are powerful to diagnose flow anomalies especially with
by moving them crosswise toward channel centerline. Accordingly, the aid of a relatively simple model [30]. This approach uses
the back side of the stretched tracer band starts displacing in the detectable ions or molecules as tracer particles which are small en-
direction of the convective flow as molecular diffusion restricts ough to be influenced by the Brownian motion of liquid to mimic
the extent of tracer convective broadening. This axial dispersion liquid diffusional behavior at the micro-scale [31].
phenomenon refers to Taylor dispersion [27]. We reexamine the A 1 mm I.D. glass capillary tube, equipped with a T-shaped
axial dispersion variations in presence of excited MNPs under injection site and two conductivity electrodes, 3-cm apart, is intro-
different magnetic field types to assess their influence on duced either vertically (for RMF and OMF tests, Fig. 2a and b), or
liquid-phase molecular transport. Axial dispersion coefficients are horizontally (for SMF, Fig. 2c) at the bore center of a vertically
measured by performing impulse residence time distribution aligned magnet. The flow of a dilute ferrofluid is maintained using
(RTD) tests in capillary. a syringe pump (Cole–ParmerÒ single-syringe infusion pump).
458 P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466

With regard to ferrofluids opaqueness even in the low concen- 2.4. Capillary tube and magnetic field relative alignment
tration range, non-intrusive optical techniques are not applicable
neither for tracer injection [32] nor for tracer detection [33,34]. The magnetic field is generated horizontally perpendicular to
Hence, we employed a conductometric technique that uses electro- magnet bore longitudinal axis. Theoretically, a dynamic magnetic
lytes (NaCl solution) at low concentration (i.e., 0.05 M) mixed with field (i.e., RMF and OMF) drag rdMNPs to spin asynchronously with
the dilute ferrofluids as a tracer. Great care was exercised in the respect to fluid vorticity [1]. Therefore, the Taylor dispersion capil-
preparation of the tracer solutions so that after dilution the MNP lary was positioned coaxially with respect to the tubular magnet in
concentration in the tracer solution must be identical to that of transverse RMF (TRMF) and transverse OMF (TOMF) experiments to
the ferrofluids flowing in the capillary. This manner prevented set the particle spin plane crosswise to the flow direction to pro-
magnetic Kelvin force [1] interference resulting from magnetic sus- mote lateral mixing (Fig. 2a and b). In contrast, SMF pins MNPs
ceptibility jump at the injection of the tracer into the capillary. At in the capillary making them to resist fluid vorticity when the mag-
t = 0, a small volume of tracer (i.e., 0.5 ll) was injected almost in- netic field direction is perpendicular to fluid vorticity (1/2rV).
stantly into the flow by applying a side channel pressure, while Since fluid vorticity lies in capillary azimuthal direction and the
cross-sectional average electrical conductivity was probed by two magnetic field is transverse to magnet longitudinal axis, the only
sets of wire electrodes 3 cm apart (Fig. 3a). The electrodes were configuration in which all MNPs oppose fluid vorticity is by adjust-
set in the capillary wall in a way to cause minimal perturbation ing the capillary horizontally and parallel to the magnetic field
of laminar flow pattern. The conductivity meter (Omega CDTX- direction (Fig. 2c).
90) generates a signal in mV reflecting the electrical conductivity
of the fluid passing by each electrode set. Holding Kohlrausch’s 2.5. Colloidal suspension
law, transient behavior of the signal intensities is associated with
electrolyte concentration change in the liquid (Fig. 3b). Dilute concentrations of ferrite (Fe3O4) MNPs (/ = 0.0005  0.01
Hydrodynamic test conditions, i.e., liquid flow rate and capillary v/v magnetic content) dispersed in water were prepared from a
diameter, d, were chosen to have small Reynolds (Re = Ud/m  O(1)) commercial ferrofluid (/ = 0.042), EMG705 (FerroTec). The mag-
and large Péclet (Pe = Ud/D  O(103), where U = 0.001 m/s, netic properties of undiluted EMG 705 were measured by an alter-
D  109 m2/s) numbers to make sure viscous forces are not dwar- nating gradient magnetometer, MicroMag model 2900 (Princeton
fed by inertial forces. However, convective transport outweighs Instrument Co.) at 298 K in low-field (for initial susceptibility,
molecular diffusion in the longitudinal direction. Each experiment v0) and high-field (for saturation magnetization, Ms) asymptote
was conducted for four Re numbers (0.50, 0.75, 1.25, and 1.50) with of magnetization curve. Using these values, particle core diameter
three repetitions for each flow rate. The axial dispersion coefficient was estimated following a method proposed by Chantrell et al.
was estimated using an axial dispersion model with open–open [36]. Table 1 summarizes the magnetic properties of undiluted
boundary conditions for laminar flow [31,35]. EMG-705 ferrofluid. Particle size distribution of dilute ferrofluid

L = 3 cm

0.08
a I II

ω
0.07 θ

m m
ω

0.06

b
Intensity

0.05

0.04 c d
I
0.03 II

0.02

0.01

0
0 20 40 60 80 100 120 140 160
Time (s)
Fig. 3. Impulse test of a Taylor dispersion capillary with and without RMF. (a) Schematic drawing of Taylor dispersion capillary tube with two sets of detectors 5 and 8 cm
down the injection point. (b) RTD responses from first (full marks) and second (empty marks) electrode. Trends represent time evolution of tracer impulse responses,
respectively, of first and second electrodes (intensity in arbitrary units): ( , ) for magnetic-field-free Poiseuille flow, ( , ) for Poiseuille flow laterally stirred by
nanomixing (/ = 0.0025, H0 = 10.4 kA/m, f = 50 Hz), ( , ) for Poiseuille flow laterally stirred by nanomixing (/ = 0.0025, H0 = 36.5 kA/m, f = 50 Hz). Peak narrowing indicates
lateral mixing under rotating magnetic field. (c and d) Schematic diagram of MNPs spin in shear flow with and without RMF. (c) In the absence of magnetic field, MNP spins,
x, gyrate collinear to fluid vorticity and MNP magnetic moments (m) are randomized in all directions. (d) Under transverse RMF (H0), x becomes perpendicular to azimuthal
fluid vorticity and hence, mixing is lateral. It is expected that a mixed zone forms around MNP when magnetic torque, m  H overcomes Brownian thermal agitation and
viscous shear forces.
P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466 459

Table 1 effect of a time-varying magnetic field [40]. A distinguishable fea-


Magnetic properties of undiluted EMG 705 from magnetometry measurement. ture of this latter system is that it is the Kelvin magnetic force
Saturation magnetization, Ms (kA/m) 18.7 which drives mixing due to MNP concentration gradients. Accord-
Initial susceptibility, v0 2.9 ingly, the mixing phenomenon per se vanishes once uniform MNP
MNP volume fraction, / (v/v) 0.042 concentration throughout the suspension is achieved. In contrast,
Estimated median magnetic core diameter, dp (nm) 16.0
the nanomixing mechanism highlighted in our study is not tied
to magnetic-force mixing effects and occurs regardless of whether
MNP concentration gradients exist or not. To distinguish such
with different concentrations was measured by magnetometry and mechanism from the micromixing mechanism which takes its
dynamic light scattering techniques (Zetasizer Nano 6, Malvern roots from the decay of fluctuating properties in turbulent flows,
Instruments Ltd.) before and after of sample exposure to magnetic we propose to christen the phenomenon observed in our (laminar
field. The results assured us of no cluster or chain formation during flow) experiments as nanomixing.
the course of experiment. Impulse RTD tests were also performed under TOMF to investi-
gate the effect of magnetic field type on excited MNP behavior at
3. Results and discussion low-Re laminar flow. For the sake of comparability, TOMF was gen-
erated using the same magnet (Fig. 2b) and imposed to the same
3.1. Taylor dispersion under TRMF and TOMF capillary configuration in identical flow condition. OMF had rms
field intensity equal to TRMF strength and same frequency. The ax-
Fig. 3b presents capillary RTD responses to impulse tests with ial dispersion coefficients for a range of liquid flow rates were esti-
and without rotating magnetic field. Magnetically-excited spinning mated from moment analysis of convoluted data [31] acquired
MNPs generate lateral mixing in capillaries as demonstrated by the from conductivity electrodes’ responses. Fig. 4 presents the capil-
attenuation of tracer band broadening relative to that in magnetic- lary axial dispersion coefficient of dilute ferrofluid (/ = 0.001) as
field-free Poiseuille flow. More symmetric trends and shorter tails dimensionless number (UL/K) versus Re number under equal
availed by two RMF excitation levels imply that diffusive migration strength of TOMF and TRMF (H0 = 31.4 kA/m) and three different
of electrolyte ions across streamlines is assisted by spinning MNPs frequencies, i.e., 10, 50 and 100 Hz. Subject to various TOMF fre-
as a secondary mixing mechanism. quencies, UL/K does not change significantly relative to magnetic
The fact that the average MNP particle–particle distance is field-free tests. This is unlike under TRMF where axial dispersion
about six times the MNP diameter (dp–p  95 nm for / = 0.0025) decreases substantially. Fig. 5 illustrates relative axial dispersion
[1], particles interaction may be neglected [37,38] and the mixing coefficient under TOMF and TRMF versus field frequency. K/K0 is al-
phenomenon can be interpreted as resulting from the individual most constant under TOMF indicating no pronounced lateral mix-
particle behavior under magnetic field. Fig. 3c shows MNPs gyrate ing effect in the capillary. In contrast, K/K0 is notably attenuated
synchronously with fluid vorticity under shear flow when there is under TRMF until a plateau is reached.
no magnetic field. Accordingly, the particle spin vector (x) is equal Figs. 4 and 5 evidence that excited MNPs under TRMF induce
to half of fluid vorticity (1/2r  v) and MNP magnetic moments lateral mixing by overcoming Brownian collisions at the molecular
(m) are randomized in all directions due to thermal agitation. level and frictional torque in the studied range of Re numbers. In
Provided TRMF yields strong enough a magnetic torque, (m  H), contrast, the very same MNPs will not give rise to lateral mixing
on individual MNPs, to overcome Brownian thermal agitation under equally strong TOMF. This observation raises the question
and hydrodynamic torque, a rotational reorientation undergone of OMF ability to bring MNPs in gyration similar to what occurs un-
by MNPs lead them to spin almost parallel to flow direction der RMF. To address this question, the MNP response mechanism
(Fig. 3d). We hypothesize that at the MNP level, a mixed fluid zone under OMF must be scrutinized further.
is formed around individual spinning MNP particles wherein Fig. 6 portrays MNP motions near capillary wall during one cy-
molecular transport phenomenon is assisted significantly with ki- cle period of transverse OMF. By exerting a torque at the very first
netic energy spread out from particle spin. part of the cycle (Fig. 6a), the magnetic field maintains the nano-
These RTD experiments in Poiseuille flows can be viewed as the particle magnetic moment almost perpendicular to both wall and
flip side of the negative-viscosity studies evidenced by ferrofluid capillary longitudinal axis. When coil electrical current and resul-
torque measurements subject to RMF [16]. Torque macroscopic tant OMF pass through zero prior to changing direction (Fig. 6b),
observations were then explicated by momentum dissipation in a a brief time lapse prevails with a magnetic field almost zero. Dur-
cylindrical cell under RMF. Likewise, spinning MNPs, while travel- ing this moment, relaxed rdMNPs are free either to rotate under
ing downstream the capillary by the pressure-driven laminar flow, the shear flow field, or to rotationally diffuse by random Brownian
dissipate kinetic energy in an anisotropic manner by preferentially collisions with the solvent molecules. The time scale, upon mag-
prompting effective lateral mixing at the nanoscale. netic field removal, during which rdMNPs get disoriented by
The lateral mixing mechanism, just evidenced above, cannot 180° from magnetic field direction due to Brownian collisions is
originate from azimuthal bulk flow of ferrofluid in TRMF (spin-up called Brownian relaxation time constant, sB [40]. It is of the order
flow). Primarily, as measured by ultrasound velocimetry [13], of 105 s for the MNPs used in our experiments [13]. Following the
spin-up flow is not observed to lead to the emergence of any radial same concept, a hydrodynamic relaxation time constant, sh, can be
velocity component which is a necessary condition for inducing defined as the time scale during which a nanoparticle is rotated by
lateral mixing in a capillary. Moreover, Khushrushahi and Zahn 180° under a shear flow field in the absence of magnetic field.
[39] recently confirmed experimentally that if RMF is truly uniform Referring to the physical view of fluid vorticity as the shear flow
over the ferrofluid volume, spin-up flow cannot arise. The fact that tendency to gyrate an infinitesimally small suspended mass, sh
our RMF is generated over a pretty narrow spatial region – a 1 mm (=2/|$  u|) amounts in average to R/(2U) = 0.25 s in our experi-
diameter capillary precisely centered at the magnet bore without ments (U = 103 m/s, R = 5  103 m). Introducing the OMF half-
any free surface-supports the hypothesis of a genuinely uniform period, sOMF, another key characteristic time coming into play
RMF with zero demagnetizing effects. It is worthy to mention that (sOMF = 0.005–0.05 s for experiments presented in Fig. 5), we have
the mixing mechanism investigated in this study occurs within a sB  sOMF < sh. Fluid vorticity is definitely too slow a mechanism to
single-phase homogeneous dilute ferrofluid. As such, it is different control MNPs rotational motion while H transitions across zero
from mixing of ferrofluids with nonmagnetic liquids under the (sOMF < sh). However, directional randomization of MNP magnetic
460 P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466

100 φ = 0.001
H0 = 31.4 kA/m

UL/K 10

Water OMF, Freq. = 100 [Hz]


Zero Field RMF, Freq. = 10 [Hz]

OMF, Freq. = 10 [Hz] RMF, Freq. = 50 [Hz]


OMF, Freq. = 50 [Hz] RMF, Freq. = 100 [Hz]
1
0.3 0.5 0.7 0.9 1.1 1.3
Re

Fig. 4. Axial dispersion coefficient in capillary with low Re number subjected to transverse OMF and RMF. Axial dispersion coefficient, compacted in dimensionless number
UL/D, estimated from impulse RTD tests. Experiments performed at / = 0.001, d/m = 1.0, H0 = 31.4 kA/m for TRMF and H0rms = 31.4 kA/m for TOMF. Axial dispersion attenuation
under TRMF implies that it changes orientation of MNP spin vectors thus inducing lateral mixing in capillary whilst under TOMF, MNP behavior does not show significant
change relative to no magnetic field tests. Error bars indicate standard deviation (triplicate tests).

K/K0 account for the fact that zero or a negative rotational viscosity un-
1.50 der OMF does not occur until the magnetic field frequency will ex-
OMF ceed MNP relaxation time constant (sBsOMF > 1) [7,9]. This is what
RMF had likely happened with the slight improvements in gas–liquid
mass transfer reported by Suresh and Bhalerao [25] and Komati
1.00 and Suresh [26] under high-frequency OMF. Recently, Sanchez
and Rinaldi [41] studied rotational viscosity of dilute ferrofluids
under oscillating and rotating magnetic fields through Brownian
dynamic simulations. In conclusion, they stated that MNPs spin
0.50
faster in RMF compared to in OMF, which explains the RMF higher
magnitude of negative rotational viscosity. Our experimental find-
φ = 0.001 ing is in agreement with this statement, as we detected a nanoscale
H0 = 31.4 kA/m
0.00 mixing effect stronger in TRMF than in TOMF.
0 20 40 60 80 100 Fig. 6 portrays part of the capillary wall on the side where TOMF
f (Hz) is perpendicular to azimuthal fluid vorticity. Complementarily,
Fig. 7 shows capillary wall at 90°-degree shifted angle from Fig. 6
Fig. 5. Relative axial dispersion coefficient in capillary under TOMF and TRMF. where TOMF is parallel to fluid vorticity. In this position, the mag-
K0 = 3.47  106 m2/s is axial dispersion coefficient of dilute ferrofluid (/ = 0.001)
netic torque does not oppose frictional torque and MNPs can rotate
without magnetic field excitation. Moderate-strength TOMF at low frequency
cannot excite MNPs to reflect in notable effects on axial dispersion. Axial dispersion freely under the shear flow field while magnetic moment direc-
attenuation under TRMF occurs and reaches a plateau after a certain frequency. tions are fully directed by TOMF. In this condition, no significant
Error bars indicate standard deviation (n = 12). momentum exchange occurs between MNPs and liquid phase
nearby the TOMF peaks. Since nanoparticle magnetic moments
are randomized in the middle of each oscillation cycle, the only
moments due to much faster Brownian agitation certainly occurs interlude where momentum exchanges are allowed between MNPs
as such reshuffling requires less time compared to OMF period and liquid is during MNP reorientation after TOMF rises from zero
(sB  sOMF). When the magnetic field strengthens again near the to its maximum strength. Such nanoparticle intermittent spin is
cycle end (Fig. 6c) it can resume an orientational coherence of isotropic and may not significantly affect lateral mixing. MNPs lie
the randomized MNP magnetic moments but in the opposite direc- between these two extreme positions conform to one of the mech-
tion. Randomness of MNPs magnetic moment directions when H anisms discussed above may not rotate continuously vis-à-vis the
passes through zero (Fig. 6b) also guarantees randomized nanopar- surrounding fluid under TOMF.
ticle spin directions during magnetic reorientation (sB  sh) hence
resulting in zero average-spin velocity per unit volume.
In RMF, superposition of three oscillating magnetic fields 3.2. Taylor dispersion under SMF
shifted by 120° out of phase (Fig. 2a) provides a resultant magnetic
field that never drops to zero. Consequently, neither Brownian col- It has been demonstrated that SMF can lock rdMNPs from viscid
lisions nor shear flow will have sufficiently wide opportunity time rotation when magnetic field is perpendicular to fluid vorticity
window to control MNP motion during a magnetic field cycle. Nev- [3,4]. Inflation of apparent viscosity in capillaries, i.e., rotational
ertheless, frequent reorientation of MNPs by TOMF and Brownian viscosity, is a well-known consequence of such MNP hindrance
agitation may generate a slight lateral mixing effect particularly [2]. Although there have been numerous studies on laminar and
for a higher-frequency field as observed on the right side of turbulent flow of ferrofluids under a coaxial SMF [10,42–49], the
Fig. 5. Due to the repeated Brownian reshuffling once in each per- transport properties of low-Re capillary flows has not received en-
iod, exposure to TOMF of MNPs is not able to sustain the synchro- ough attention to date. Thus, we performed Taylor dispersion tests
nized parallel spin seen with TRMF (Fig. 3d). This phenomenon may in capillaries to clarify in which manner interactions under SMF
P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466 461

Fig. 6. MNP response to TOMF in a position where magnetic field direction is perpendicular to fluid vorticity, (a) magnetic field is strong enough to orient MNPs to its
dominant direction, (b) TOMF passes briefly through zero while changing direction leaving MNP magnetic moments shortly unassisted to lose, due to randomization effect of
Brownian collisions, their coherent direction, (c) magnetic field on the rise until peak and MNP magnetic moments to resume orientational coherence. Synchronous particle
rotation driven by TOMF is prohibited by midway Brownian reshuffling of MNPs.

between excited MNPs and laminar flow fields may influence lat- ited by the pinned MNPs which reflects in inflated axial dispersions
eral mass transfer. under ASMF. Stokes–Einstein law for diffusion in solutions (Eq. (2))
As discussed earlier in reference with Fig. 7, excited MNPs are relates molecular diffusivity (D) to the reciprocal of liquid dynamic
unable to exchange momentum with the flow if their magnetic viscosity (g) [50]. Viewing diffusion through Stokes–Einstein law;
moments are aligned with fluid vorticity. To prevent this instance we expect that apparent molecular diffusivity reduces as apparent
from happening, the capillary tube was positioned horizontally in viscosity augments with ASMF.
the magnet where the magnetic field is coaxial with the capillary
longitudinal axis (Fig. 2c). In this configuration, the magnetic field kB T
is perpendicular to the azimuthal fluid vorticity over the entire D¼ ð2Þ
6pg0 Rp
capillary cross-section.
Fig. 8 shows RTD impulse responses with ( , ) and without ( , To explain this observation, we propose a mechanism to de-
) axial SMF (ASMF). Signals from upstream electrodes ( , ) are scribe changes in velocity profile and tracer concentration gradient
very similar whilst, downstream signals ( , ) show marked devi- at nanometric scale in the presence of excited MNPs. Magnetic tor-
ations from each other. The resemblance of inlet signals in pres- que locks MNPs to prevent their rotation under viscous torque. The
ence and absence of ASMF can be attributed to the short distance locked particles prevent liquid layers of laminar flow from sliding
(1 cm) between tracer injection and first electrode. Therefore, the on top of each other, as compared to those in the absence of mag-
tracer residence time before reaching the first electrode is not suf- netic field (Fig. 9). Thus, laminar velocity profile may change from a
ficient to reflect the effect of excited MNPs on axial dispersion. The smooth parabolic shape to a rugged form in nanoscale with some
1st-electrode skewed tracer responses are ascribed to imperfect local zero-velocity-gradient segments (Fig. 9b). In Taylor disper-
tracer injections that are well accounted for in the convolution sion test, radial tracer concentration gradient is induced originally
integral for the estimation of axial diffusion coefficients. The outlet by velocity gradient. This concentration gradient may drop to zero
signal under ASMF ( ) is wider and its breakthrough time occurs in certain positions over several contiguous stream layers wherein
earlier than without magnetic field (Fig. 8). Unlike in TRMF, this axial velocity is almost uniform. Consequently, radial mass transfer
observation demonstrates that axial dispersion has been inflated due to molecular diffusion declines over these narrow regions, as
even further by ASMF-excited MNPs interacting with shear flow. there is less driving force when some neighboring streamlines
Axial dispersion is caused by the stretching effect of convective are pinned by excited MNPs. Compensation imposed by the conti-
flow while molecular diffusion restricts it by ensuring lateral trans- nuity equation suggests there must be other segments where lin-
port over capillary cross-section [27]. At the same liquid flow rate ear velocity jumps between liquid layers (high velocity gradient)
lateral molecular diffusion of tracer appears to have been prohib- result in large tracer concentration gradient (Fig. 9b). However,
462 P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466

MNPs as conjectured in this study. The plateauing trend observed


with increasing MNP concentration in ASMF (Fig. 10b) may imply
two features. First, it confirms that MNPs are unlikely to obstruct
tracer diffusion bodily, otherwise that would have resulted in stea-
dily increasing trend of K/K0 when MNPs interparticle distance
shortens by factor 1.5 (0.0025 < / < 0.01). Second, the plateau sug-
gests that uniform velocity segments grow among non-uniform
segments to an extent whereby hydrodynamic torque dominates
magnetic torque and controls MNP gyration in slimed steepened
non-uniform segments.
Ferrofluid positive rotational viscosity under ASMF in a tube has
been reported in the literature from rheometry measurement [3,4].
In this study, we measured lesser lateral mass transfer rate or
equivalently inferior apparent liquid diffusivity in capillary as com-
pared to that driven solely by molecular diffusion in magnetic
field-free tests. These results together with those under TRMF sug-
gest that extension of Stokes–Einstein law (Eq. (2)) to apparent vis-
cosity and apparent diffusivity of ferrofluid under TRMF and ASMF
is valid qualitatively. Many applications in microfluidics require a
tool to control axial dispersion coefficient in micro-channels [52],
such as reactant homogenization in chemical reactions [53] and
protein folding [54].

3.3. Excited MNPs alter laminar velocity profile in capillary flow

We further investigated how excited MNPs may affect the par-


abolic velocity profile of a laminar Poiseuille flow. We expect that
axially spinning MNPs as depicted in Fig. 3d for TRMF attenuate ra-
dial velocity gradients and result in more flattened velocity profiles
compared to magnetic-field-free parabolic shape. Locked MNPs
under ASMF in contrast augment radial velocity gradient resulting
in a magneto-thickening fluid behavior.
Figs. 3b and 8b provide a crucial clue as the time lag between
breakthrough times changes with nanomixing and nano-hindrance
despite an identical mean residence time (t ¼ L=U) for both tests.
Breakthrough time features the fastest centerline tracer blob pass-
ing by the electrodes with minimum residence time (tmin). Hence,
Fig. 7. MNP response to OMF in a position where magnetic field direction is parallel
t=t min ratio can be used as a quantitative index to infer how veloc-
to fluid vorticity. MNP gyration under shear flow is not opposed by magnetic
torque. As a result, there is no momentum transfer between MNP and fluid due to ity profiles change due to the presence of excited MNPs. This can be
OMF–nanoparticle interactions. accomplished by drawing an analogy with non-Newtonian power-
law radial velocity profiles, u(r) [55]:
  r 1þ1=n 
faster transfer rate over these segments may not compensate for 3n þ 1
uðrÞ ¼ U 1 ð3Þ
the global mass transfer decline; since molecular transport be- 1þn R
tween streamlines occurs in-series and thus overall mass transfer
rate is limited by the slowest slice. A consequence to this view dic- where the power-law index, n, is in the range of 0 (flat profile) to 1
tated by continuity equation is that the maximum linear velocity (parabolic profile) and expressed as a function of tmin =t according to
on capillary’s longitudinal axis must be higher than that in mag- [55]:
netic-free condition (Fig. 9b). This evidence will be thoroughly ana-
1  t min =t
lyzed later based on minimum (or breakthrough) residence time n¼ ð4Þ
3tmin t  1
(tmin) determinations.
To gain more insights into the overall diffusion inhibition, we Power-law index, n, declines from 1 to 0 for a velocity profile
have further compared K dependency on H0 (0–35 kA/m) and / evolving from parabolic to flat but increases over 1 for velocity pro-
(0–0.01) under TRMF and ASMF. K/K0 is plotted as a function of files more fusiform than parabolic.
H0 (Fig. 10a) and / (Fig. 10b) where K0 refers to axial dispersion Fig. 11a and b plots t=t min versus H0 and / for Fig. 10a and b
coefficient at H0 = 0 (Fig. 10a) and / = 0 (Fig. 10b). Since axial spin- experiments. Magnetic field strength and particle concentration
ning or hindrance of MNPs is magnetic-torque driven, K/K0 rises promote flattening and protruding effects of excited MNPs on lam-
under ASMF and its decline under TRMF is intensified by H0 magni- inar velocity profile. Plateaued trends of Fig. 11b suggest that
tude. The other factor promoting both phenomena was found to be occurrence of local uniform velocity segments due to inter-stream-
MNP concentration (Fig. 10b). For magnetic core volume fractions line nanometric radial momentum transfer under ASMF will be
ranging from 103 to 102, the average MNP interparticle distance limited by wall-induced shear stress. This mechanism prevents
varies from 129 nm to 60 nm for dp = 16.0 nm [1]. At such dilution uniform segments in capillary to grow and merge to lead to a
levels, the relative vortex and shear viscosities are close to 0 and 1, whole flattened velocity profile (Fig. 9b). Hence, lateral momentum
respectively [51], and magnetic interactions between rigid dipole transfer in capillary in terms of homogenization of linear velocity
MNPs may be neglected. At such concentration levels, particle-flow may not be assisted effectively by excited MNPs except under
T
field interaction may result only from the behavior of individual RMF. MNP-mediated hydrodynamic effects on velocity profile is
P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466 463

Fig. 8. Impulse test of a Taylor dispersion capillary with and without SMF. (a) Schematic drawing of Taylor dispersion capillary tube with two sets of detectors one and four
cm down the injection point. (b) RTD responses from first (full marks) and second electrode (empty marks). Trends represent the time evolution of tracer intensity from first
and second electrodes, respectively, ( , ) with unexcited MNPs flowing in laminar Poiseuille flow, ( , ) with excited MNPs (/ = 0.0025, H0 = 31.4 kA/m). Axial dispersion
increases in capillary when magnetic torque prevents MNPs gyration under shear flow.

a K/K 0
10.0
SMF
RMF

φDCF = 0.005
φRMF = 0.001
1.0 fRMF = 50 Hz

0.1
0.0 1.0 2.0 3.0 4.0
H0×10 4(A/m)
K/K0
b 10.0
SMF
RMF

1.0 H0 = 31.4 kA/m


fRMF = 50 Hz
Fig. 9. Schematic of MNPs motion under shear flow in capillary tube. (a) Magnetic
field free, particles gyrate by friction torque and their magnetic moments directions
are randomized. (b) Under ASMF, hard dipoles are locked by the coaxially applied 0.1
external magnetic field. Nanometrically, the velocity gradient has been removed in
some regions (where highlighted) and augmented in the rest to satisfy continuity
(black dotted line). That may result in mimetic shear thickening behavior of liquid
under ASMF. The brown dash-dotted line represents the original parabolic velocity 0.0
profile as it occurs in (a). 0.0 0.2 0.4 0.6 0.8 1.0
φ×10 -2

Fig. 10. Axial dispersion perturbations under TRMF and ASMF versus (a) magnetic
illustrated in Fig. 11c which shows computed, flattened or pro- field strength and (b) MNP concentration. (a and b) ASMF promotes axial dispersion
truded, velocity profiles as t=tmin ratio deviate from 2 to echo nano- ( , ) whilst TRMF attenuates K/K0 ( , ). Error bars indicate standard deviation
mixing or nano-hindrance intensities. (number of repeat runs = 6). Magnetically locked MNPs under ASMF reduce lateral
mass transfer rate in capillary whereas magnetically spinning MNPs under TRMF
Except ultrasound velocimetry [13,39,47,56], no other tech-
promote lateral mass transfer through nanoconvective mixing. In (a) K0(/ = 0.001,
nique has been proposed in the literature to measure ferrofluid T
RMF) = 3.47  106 m2/s and K0(/ = 0.005, ASMF) = 1.13  106 m2/s. In (b) K0(/
velocity profiles. This technique is not reliable in regions closer = 0, TRMF) = 5.08  106 m2/s and K0(/ = 0, ASMF) = 1.11  106 m2/s.
than several millimeters to the wall due to echo interference
[57,58] and therefore, may not be applicable in 1 mm micro-chan-
nel as well. Furthermore, velocity profile reconstruction confirmed ferrofluid under ASMF and TRMF in capillary. Although both effects
magneto-thickening and magneto-thinning behavior of dilute would be perceived from positive [3,4,15,16] or negative
464 P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466

These two equations assume that lateral (radial) transfer of


a t/tmin
solute depends only on radial variation of tracer concentration
SMF φSMF = 0.005
4.0 RMF φRMF = 0.001 and as such mass transfer is only driven by molecular diffusion.
fRMF = 50 Hz Thus, D appears in the denominator of Eqs. (5) and (6). This
assumption is violated in our study as the nanoconvective effect
3.0 by MNPs enhances or retards lateral mixing. Keeping other
assumptions of the model still valid, the molecular diffusivity must
be replaced by an effective diffusivity (Deff) in the modified Taylor
model of axial dispersion:
2.0

R2 U 2 n2
K¼ ð7Þ
Deff 2ð3n þ 1Þð5n þ 1Þ
1.0
0.0 1.0 2.0 3.0 4.0 For a parabolic profile (n = 1) and no magnetic field effects, Eq.
H0×104 (A/m) (7) becomes:

t/tmin R2 U 2
b SMF K0 ¼
48D
ð8Þ
3.0 RMF
Thus,

Deff K 0 48n2
¼ ð9Þ
D K 2ð3n þ 1Þð5n þ 1Þ
H0 = 31.4 kA/m
2.0 fRMF = 50 Hz Fig. 12 is a plot of Deff/D as a function of magnetic field strength
(Fig. 12a) and MNP volume fraction (Fig. 12b). The Deff/D ratios
were estimated from the K/K0 values shown in Fig. 10a and b and
their corresponding t=t min values of Fig. 11a and b. In agreement
with previous discussion of Figs. 10 and 11, Fig. 12 demonstrates
1.0 a systematic augmentation of effective diffusivity by up to a factor
0.0 0.1 0.2 0.3 0.4 0.5
three in the presence of spinning MNPs in rotating magnetic fields.
φ×10 -2 The same factor also reflects in terms of reduction of effective dif-
u /U fusivity under a static magnetic field. Hence magnetically pinned
c 4.0 MNPs even incapacitated molecular diffusivity in smoothing the
t/tmin = 4.0 n =-3.0
3.5 t/tmin = 3.0 n = Inf
t/tmin = 2.5 n = 3.0
Deff/D
3.0
t/tmin = 2.0 n = 1.0 a 2.0
2.5 t/tmin = 1.2 n = 0.1 SMF
t/tmin = 1.0 n = 0.0 RMF
2.0
1.5
1.5 φDCF = 0.005
φRMF = 0.001
1.0 1.0 fRMF = 50 Hz

0.5

0.0 0.5
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5
r (mm)
0.0
Fig. 11. Excited MNPs effect on laminar velocity profile under TRMF and ASMF. 0.0 1.0 2.0 3.0 4.0
Power-law index n and t=tmin ratio as a function of (a) magnetic field strength, (b) H0×104 (A/m)
MNP volume fraction. Error bars indicate standard deviation (number of repeat
Deff/D
runs = 6). (c) Expected laminar flow velocity profile in a capillary as a function of
power-law index.
b 4.0
SMF
RMF
[7,9,15,16] rotational viscosity, they have never been spotted 3.0
experimentally from the perspective of manipulating laminar
velocity profiles in capillaries.
Taylor [28] expressed the dispersion coefficient (K) of a solute 2.0 H0 = 31.4 kA/m
flowing slowly through a tube in terms of U, d and D for a parabolic fRMF = 50 Hz
velocity profile as:
1.0
R2 U 2
K¼ ð5Þ
48D
0.0
Later, this equation was adapted and extended to a power-law 0.0 0.2 0.4 0.6 0.8 1.0
velocity profile 60:
φ×10-2
R2 U 2 n2
K¼ ð6Þ Fig. 12. Effective diffusivity in laminar flow under TRMF and ASMF. Deff/D ratio as a
D 2ð3n þ 1Þð5n þ 1Þ function of (a) magnetic field strength and (b) MNP volume fraction.
P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466 465

radial concentration gradient in capillary. Reduction of micro-scale [11] F. Gazeau, C. Baravian, J.C. Bacri, R. Perzynski, M.I. Shliomis, Energy conversion
in ferrofluids: magnetic nanoparticles as motors or generators, Phys. Rev. E. 56
radial concentration gradient around pinned MNPs has been ac-
(1997) 614–618.
counted for Deff reduction as explained with respect to Fig. 9. [12] R. Moskowitz, R.E. Rosensweig, Nonmechanical torque-driven flow of a
ferromagnetic fluid by an electromagnetic field, Appl. Phys. Lett. 11 (1967)
301–303.
4. Conclusion [13] A. Chaves, C. Rinaldi, S. Elborai, X. He, M. Zahn, Bulk flow in ferrofluids in a
uniform rotating magnetic field, Phys. Rev. Lett. 96 (2006) 194501.
[14] A. Chaves, M. Zahn, C. Rinaldi, Spin-up flow of ferrofluids: asymptotic theory
Experimental works to elucidate the effect of magnetically-ex- and experimental measurements, Phys. Fluids. 20 (2008) 053102.
cited MNPs on molecular transport mechanism in ferrofluids is [15] A.D. Rosenthal, C. Rinaldi, T. Franklin, M. Zahn, Torque measurements in spin-
up flow of ferrofluids, J. Fluid Eng-Trans ASME 126 (2004) 198–205.
lacking. By studying axial dispersion in dilute ferrofluids flowing [16] C. Rinaldi, F. Gutman, X.W. He, A.D. Rosenthal, M. Zahn, Torque measurements
through a Taylor capillary cell under moderate strength and low on ferrofluid cylinders in rotating magnetic fields, J. Magn. Magn. Mater. 289
frequency TRMF, TOMF and ASMF, we investigated the mass trans- (2005) 307–310.
[17] L.D. Mao, S. Elborai, X.W. He, M. Zahn, H. Koser, Direct observation of closed-
fer phenomena induced by the interactions of excited MNPs with loop ferrohydrodynamic pumping under traveling magnetic fields, Phys. Rev. B
Poiseuille flow fields. Although any type of magnetic field gives rise 84 (2011) 104431.
to a bodily magnetic torque on rdMNPs, our study demonstrates [18] T.H. Chilton, A.P. Colburn, Mass transfer (absorption) coefficients – prediction
from data on great transfer and fluid friction, Ind. Eng. Chem. 26 (1934) 1183–
that the nature of the imposed magnetic field must be accounted
1187.
for as a crucial factor in determining the particle–field interactions. [19] S. Shylesh, V. Schunemann, W.R. Thiel, Magnetically separable nanocatalysts:
The experimental results point out that rdMNPs in TRMF and ASMF bridges between homogeneous and heterogeneous catalysis, Angew. Chem.
manipulate tracer lateral diffusion rate in capillary whilst particle Int. Ed. 49 (2010) 3428–3459.
[20] P.D. Stevens, G.F. Li, J.D. Fan, M. Yen, Y. Gao, Recycling of homogeneous Pd
excitation by TOMF has no significant impact on it. catalysts using superparamagnetic nanoparticles as novel soluble supports for
Furthermore, we correlated the shape of ferrofluid laminar Suzuki, Heck, and Sonogashira cross-coupling reactions, Chem. Commun.
velocity profile in capillary to the breakthrough (minimum) resi- (2005) 4435–4437.
[21] M. Shokouhimehr, Y. Piao, J. Kim, Y. Jang, T. Hyeon, A magnetically recyclable
dence time from impulse RTD test data. Using a single parameter nanocomposite catalyst for olefin epoxidation, Angew. Chem. Int. Ed. 46
model in this technique, we evidenced that ferrofluid laminar (2007) 7039–7043.
velocity profiles may change in ASMF or TRMF. More investigations [22] C. Che, W. Li, S. Lin, J. Chen, J. Zheng, J.-c. Wu, Q. Zheng, G. Zhang, Z. Yang, B.
Jiang, Magnetic nanoparticle-supported Hoveyda–Grubbs catalysts for ring-
are required on the shape of velocity profile by an optical veloci- closing metathesis reactions, Chem. Commun. (2009) 5990–5992.
metry technique to verify the reported results. [23] X.X. Zheng, S.Z. Luo, L. Zhang, J.P. Cheng, Magnetic nanoparticle supported
In conclusion, we found that rdMNP rotational motion relative ionic liquid catalysts for CO(2) cycloaddition reactions, Green Chem. 11 (2009)
455–458.
to contiguous liquid serves as a valuable molecular mixing tool [24] V. Polshettiwar, R. Luque, A. Fihri, H. Zhu, M. Bouhrara, J.-M. Bassett,
for intensification or retardation of diffusion in liquids, especially Magnetically recoverable nanocatalysts, Chem. Rev. 111 (2011) 3036–3075.
in sub-micronic thin regions. Control of the dispersion phenome- [25] A.K. Suresh, S. Bhalerao, Rate intensification of mass transfer process using
ferrofluids, Indian J. Pure Appl. Phys. 40 (2001) 172–181.
non in a microfluidic context may be considered as a prime exam-
[26] S. Komati, A.K. Suresh, CO2 absorption into amine solutions: a novel strategy
ple. The fact that suspended functionalized MNPs find extensive for intensification based on the addition of ferrofluids, J. Chem. Technol.
applications in a variety of disciplines [59–61] will broaden the Biotechnol. 83 (2008) 1094–1100.
scope of potential application of this technique. Moreover, this ap- [27] H. Brenner, D.A. Edwards, Macrotransport Processes, Butterworth-Heinemann,
Boston, 1993.
proach opens up a new trend in experimental methods to shed [28] G. Taylor, Dispersion of soluble matter in solvent flowing slowly through a
light on the interaction between magnetically excited MNPs and tube, Proc. R. Soc. London, Ser. A 219 (1953) 186–203.
flow field at the nanoscale as shown in our recent applications of [29] G.I. Taylor, Diffusion and mass transport in tubes, Proc. Phys. Soc. B 67 (1954)
857–869.
this technique [62–65]. [30] O. Levenspiel, Modeling in chemical engineering, Chem. Eng. Sci. 57 (2002)
4691–4696.
[31] O. Levenspiel, Chemical Reaction Engineering, John Wiley & Sons, New York,
Acknowledgements 1999.
[32] 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
Support from the Natural Sciences and Engineering Research microreactors, Lab Chip 8 (2008) 431–438.
Council of Canada and the Canada Research Chair ‘‘Green processes [33] F. Trachsel, A. Gunther, S. Khan, K.F. Jensen, Measurement of residence time
for cleaner and sustainable energy’’ is gratefully acknowledged. distribution in microfluidic systems, Chem. Eng. Sci. 60 (2005) 5729–5737.
[34] M. Gunther, S. Schneider, J. Wagner, R. Gorges, T. Henkel, M. Kielpinski, J.
The authors also thank A. Faridkhou and O. Gravel for helpful edit- Albert, R. Bierbaum, J.M. Kohler, Characterisation of residence time and
ing assistance. residence time distribution in chip reactors with modular arrangements by
integrated optical detection, Chem. Eng. J. 101 (2004) 373–378.
[35] O. Levenspiel, W.K. Smith, Notes on the diffusion-type model for the
References longitudinal mixing of fluids in flow, Chem. Eng. Sci. 6 (1957) 227–233.
[36] R.W. Chantrell, J. Popplewell, S.W. Charles, Measurements of particle-size
distribution parameters in ferrofluids, IEEE Trans. Magn. 14 (1978) 975–977.
[1] R.E. Rosensweig, Ferrohydrodynamics, Dover Publications, Mineola, N.Y., 1997.
[37] S.P. Gubin, Y.A. Koksharov, G.B. Khomutov, G.Y. Yurkov, Magnetic
[2] S. Odenbach, Magnetoviscous effects in ferrofluids, Springer, Berlin, 2002.
nanoparticles: preparation methods, structure and properties, Usp. Khim. 74
[3] J.P. McTague, Magnetoviscosity of magnetic colloids, J. Chem. Phys. 51 (1969)
(2005) 539–574.
133.
[38] H.T. Yang, H.L. Liu, N.N. Song, H.F. Du, X.Q. Zhang, Z.H. Cheng, J. Shen, L.F. Li,
[4] R.E. Rosensweig, R. Kaiser, G. Miskolcz, Viscosity of magnetic fluid in a
Determination of the critical interspacing for the noninteracting magnetic
magnetic field, J. Colloid Interface Sci. 29 (1969) 680–686.
nanoparticle system, Appl. Phys. Lett. 98 (2011) 153112.
[5] S. Rhodes, X.W. He, S. Elborai, S.H. Lee, M. Zahn, Magnetic fluid behavior in
[39] S. Khushrushahi, M. Zahn, Ultrasound velocimetry of ferrofluid spin-up flow
uniform DC, AC, and rotating magnetic fields, J. Electrostat. 64 (2006) 513–519.
measurements using a spherical coil assembly to impose a uniform rotating
[6] M.I. Shliomis, Effective viscosity of magnetic suspensions, Sov. Phys. JETP-USSR
magnetic field, J. Magn. Magn. Mater. 323 (2011) 1302–1308.
34 (1972) 1291.
[40] L. Mao, H. Koser, Overcoming the diffusion barrier: ultra-fast micro-scale
[7] J.C. Bacri, R. Perzynski, M.I. Shliomis, G.I. Burde, Negative-viscosity effect in a
mixing via ferrofluids, in: Proceedings of the 14th International Conference on
magnetic fluid, Phys. Rev. Lett. 75 (1995) 2128–2131.
Solid-State Sensors, Actuators and Microsystems, IEEE, Lyon, 2007, p. 1829.
[8] M.I. Shliomis, K.I. Morozov, Negative viscosity of ferrofluid under alternating
[41] A.H. Morrish, The Physical Principles of Magnetism, IEEE Press, New York,
magnetic-field, Phys. Fluids 6 (1994) 2855–2861.
2001.
[9] A. Zeuner, R. Richter, I. Rehberg, Experiments on negative and positive
[42] J.H. Sanchez, C. Rinaldi, Magnetoviscosity of dilute magnetic fluids in
magnetoviscosity in an alternating magnetic field, Phys. Rev. E. 58 (1998)
oscillating and rotating magnetic fields, Phys. Fluids 22 (2010) 043304.
6287–6293.
[43] R. Patel, R. Upadhyay, R.V. Mehta, Viscosity measurements of a ferrofluid:
[10] K.R. Schumacher, I. Sellien, G.S. Knoke, T. Cader, B.A. Finlayson, Experiment and
comparison with various hydrodynamic equations, J. Colloid Interface Sci. 263
simulation of laminar and turbulent ferrofluid pipe flow in an oscillating
(2003) 661–664.
magnetic field, Phys. Rev. E. 67 (2003) 026308.
466 P. Hajiani, F. Larachi / Chemical Engineering Journal 223 (2013) 454–466

[44] N. Andhariya, B. Chudasama, R. Patel, R.V. Upadhyay, R.V. Mehta, Field induced [56] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, J. Wiley, New
rotational viscosity of ferrofluid: effect of capillary size and magnetic field York, 2002. p. 232.
direction, J. Colloid Interface Sci. 323 (2008) 153–157. [57] I. Torres-Díaz, C. Rinaldi, S. Khushrushahi, M. Zahn, Observations of ferrofluid
[45] L.M. Pop, S. Odenbach, Capillary viscosimetry on ferrofluids, J. Phys.: Condens. flow under a uniform rotating magnetic field in a spherical cavity, J. Appl. Phys.
Matter 20 (2008) 204139. 111 (2012) 07B313.
[46] M. Reindl, A. Leschhorn, M. Luecke, S. Odenbach, in: H. Yamaguchi (Ed.), 12th [58] I. Torres-Diaz, C. Rinaldi, Ferrofluid flow in the annular gap of a multipole
International Conference on Magnetic Fluids Icmf12, vol. 9, 2010, p. 121. rotating magnetic field, Phys. Fluids 23 (2011) 082001.
[47] M. Reindl, S. Odenbach, Effect of axial and transverse magnetic fields on the [59] A. Debrassi, A.F. Correa, T. Baccarin, N. Nedelko, A. Slawska-Waniewska, K.
flow behavior of ferrofluids featuring different levels of interparticle Sobczak, P. Dluzewski, J.-M. Greneche, C.A. Rodrigues, Removal of cationic dyes
interaction, Phys. Fluids 23 (2011) 093102. from aqueous solutions using N-benzyl-O-carboxymethylchitosan magnetic
[48] M. Reindl, S. Odenbach, Influence of a homogeneous axial magnetic field on nanoparticles, Chem. Eng. J. 183 (2012) 284–293.
Taylor–Couette flow of ferrofluids with low particle–particle interaction, Exp. [60] Y. Pang, G. Zeng, L. Tang, Y. Zhang, Y. Liu, X. Lei, Z. Li, J. Zhang, Z. Liu, Y. Xiong,
Fluids 50 (2011) 375–384. Preparation and application of stability enhanced magnetic nanoparticles for
[49] K.R. Schumacher, J.J. Riley, B.A. Finlayson, Effects of an oscillating magnetic rapid removal of Cr(VI), Chem. Eng. J. 175 (2011) 222–227.
field on homogeneous ferrofluid turbulence, Phys. Rev. E. 81 (2010) 016317. [61] Z.-p. Yang, X.-y. Gong, C.-j. Zhang, Recyclable Fe(3)O(4)/hydroxyapatite
[50] K.R. Schumacher, J.J. Riley, B.A. Finlayson, Turbulence in ferrofluids in channel composite nanoparticles for photocatalytic applications, Chem. Eng. J. 165
flow with steady and oscillating magnetic fields, Phys. Rev. E. 83 (2011) (2010) 117–121.
016307. [62] A. Pohar, P. Žnidaršič-Plazl, I. Plazl, Integrated system of a microbioreactor and
[51] C.C. Miller, The Stokes Einstein law for diffusion in solution, Proc. R. Soc. Lond. a miniaturized continuous separator for enzyme catalyzed reactions, Chem.
A 106 (1924) 724–749. Eng. J. 189–190 (2012) 376–382.
[52] S. Feng, A.L. Graham, J.R. Abbott, H. Brenner, Antisymmetric stresses in [63] P. Hajiani, F. Larachi, Giant liquid-self diffusion in stagnant liquids by magnetic
suspensions: vortex viscosity and energy dissipation, J. Fluid Mech. 563 (2006) nanomixing, Chem. Eng. Process. Process Intensif., in press, doi: 10.1016/
97–122. j.cep.2013.01.014.
[53] S.W. Jones, W.R. Young, Shear dispersion and anomalous diffusion by chaotic [64] P. Hajiani, F. Larachi, Remotely excited magnetic nanoparticles and gas-liquid
advection, J. Fluid Mech. 280 (1994) 149–172. mass transfer in Taylor flow regime, Chem. Eng. Sci., in press, doi: 10.1016/
[54] A.J. deMello, Control and detection of chemical reactions in microfluidic j.ces.2013.01.052.
systems, Nature 442 (2006) 394. [65] P. Hajiani, F. Larachi, Reducing Taylor dispersion in capillary laminar flows
[55] S.S. Varghese, Y.G. Zhu, T.J. Davis, S.C. Trowell, FRET for lab-on-a-chip devices – using magnetically excited nanoparticles: Nanomixing mechanism for micro/
current trends and future prospects, Lab Chip 10 (2010) 1355–1364. nanoscale applications, Chem. Eng. J. 203 (2012) 492–498.

You might also like