Você está na página 1de 9

Please cite this article in press as: A. Bahmanyar, et al.

, The inuence of nanoparticles on hydrodynamic characteristics and mass transfer


performance in a pulsed liquidliquid extraction column. Chem. Eng. Process. (2011), doi:10.1016/j.cep.2011.08.008
ARTICLE IN PRESS
GModel
CEP-6047; No. of Pages 9
Chemical Engineering and Processing xxx (2011) xxxxxx
Contents lists available at SciVerse ScienceDirect
Chemical Engineering and Processing:
Process Intensication
j our nal homepage: www. el sevi er . com/ l ocat e/ cep
The inuence of nanoparticles on hydrodynamic characteristics and mass
transfer performance in a pulsed liquidliquid extraction column
Amir Bahmanyar
a
, Naseh Khoobi
b
, Mohammad Reza Mozdianfard
a
, Hossein Bahmanyar
c,
a
Separation Processes Research Group (SPRG), Department of Engineering, University of Kashan, Iran
b
Department of Chemical Engineering, Sharif University of Technology, Tehran, Iran
c
Surface Phenomena and Liquid-Liquid Extraction Research Laboratory, University of Tehran, Tehran, Iran
a r t i c l e i n f o
Article history:
Received 28 April 2011
Received in revised form4 August 2011
Accepted 16 August 2011
Available online xxx
Keywords:
Pulsed column
Dynamic hold-up
Static hold-up
Enhancement factor
Nanouids
a b s t r a c t
With respect to the inuence of nanoparticles on mass transfer characteristics, limited number of studies
available in the literature, deal primarily with gasliquid systems. In this work, mass transfer perfor-
mance and hydrodynamic characteristics including static and dynamic dispersed phase hold-ups of
nanouids have been investigated for pulsed liquidliquid extraction column (PLLEC). The nanouids
used were prepared by dispersing SiO
2
nanoparticles of 0.01, 0.05 and 0.1 volume percent with two
different hydrophobicities in kerosene as base uid using ultrasonication. UVvis spectrophotometer
was also used for evaluation of the nanouids stability. The results were compared with conditions of
no-nanoparticles in the dispersed phase and in the absence of mass transfer effect (no acetic acid as
solute). Different pulsation intensities were maintained for the xed mass ow rates of dispersed (Q
d
)
and continuous (Q
c
) phases (with ratio Q
c
/Q
d
= 1.2) with mass transfer direction being made from the
dispersed phase to the continuous one. The results indicate that in the presence of the nanouids, static
and dynamic dispersed phase hold-ups increased by 23398%, and 23257%, respectively, while mass
transfer performance was enhanced by 460%.
2011 Elsevier B.V. All rights reserved.
1. Introduction
Nanouids have been extensively studied within the last two
decades, primarily in association with their heat transfer enhance-
ment characteristics. Hence, classic metal conductive nanoparticles
such as Al and Cu [1] and their oxides [2] have been used with base
uids including water and ethylene glycol which have increased
thermal conductivity by 1540%. Some studies [3] have attempted
measurement of convective heat transfer coefcient, claimed to be
far moredifcult thanconductiveone[4] andreported60%increase
in the heat transfer coefcient [5]. These studies which are mostly
conducted in the laboratory scale have employed nanoparticles of
less than 100nm, 0.54 volume percent (vol%) and temperature
range of 2050

C.
The inuences of nanoparticles on heat transfer are reviewed
adequately in the literature [4,6]. However, by comparison, the
inuence of nanoparticles onmass transfer operations has received
much less attention. This approach has been initiated by con-
sideration of Brownian movement of the nanoparticles by some
researchers to be responsible as the primary factor in the enhance-

Corresponding author. Tel.: +98 21 61112213.


E-mail address: hbahmany@ut.ac.ir (H. Bahmanyar).
ment characteristics of nanouids in both convective [7] and
conductive [8] heat transfer.
Assuming convection and mass transfer to be similar in nature,
Krishnamurthy et al. [9] studied mass diffusion of uoresein dye in
nanouids by taking time-dependent images. Their images, illus-
trated much faster dye diffusion in nanouids than in pure water
and their calculations of mean displacement equation suggested
that the Brownian motion of the nanoparticles did not contribute
directly tothe mass transport enhancement; rather it is the velocity
disturbance eld in the uid, created by the motion of the nanopar-
ticles which is most likely responsible for such enhancement. In
other words, the enhancement in diffusivity is likely due to the
increased nanoscale stirring of the liquid, caused by the nanopar-
ticles Brownian motion. They also observed an unexpected clear
peak in the enhanced diffusivity of the dye in nanouid which they
could not justify.
In several previous studies [10,11], the effect of particles in
micron size have been investigated on bubble columns and adsorp-
tionoperations. Signicant enhancement of gas adsorption(usually
CO
2
) in various liquids has been attributed to grazing or shut-
tle mechanism [12] believed to be the transfer phenomenon
mechanism responsible for a gas from the gasliquid interface to
the bulk of the liquid [13]. With recent advancement in nano-
technology, previous micro-sized applications are being extended
to nanoparticles, and following meeting challenges such as
0255-2701/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cep.2011.08.008
Please cite this article in press as: A. Bahmanyar, et al., The inuence of nanoparticles on hydrodynamic characteristics and mass transfer
performance in a pulsed liquidliquid extraction column. Chem. Eng. Process. (2011), doi:10.1016/j.cep.2011.08.008
ARTICLE IN PRESS
GModel
CEP-6047; No. of Pages 9
2 A. Bahmanyar et al. / Chemical Engineering and Processing xxx (2011) xxxxxx
Fig. 1. Schematic of the pulsed liquidliquid extraction column used in this study.
stability of phase containing nanoparticles and agglomeration
diverse possible applications of nanoparticles in mass transfer
operations are being opened up. Falling lmowof binary nanou-
ids were studied by Kang et al. [13] who measured the vapor
absorption rate as well as heat transfer rate for falling lmow of
binary nanouids. They used H
2
O/LiBr solution with nanoparticles
of FeandCarbonnanotubes (CNT) withvarious concentrations of up
to 0.1wt% as their binary nanouids and found that vapor absorp-
tion rate increased with increasing the solution mass owrate and
the concentration of Fe and CNT nanoparticles. More importantly,
they illustrated that mass transfer enhancement due to nanoparti-
cles presence was much more signicant than that of heat transfer.
Oxygen-transfer enhancement in the presence of colloidal disper-
sions of magnetite (Fe
3
O
4
) nanoparticles coated with oleic acid is
also observed by Olle et al. [14] who claimed gasliquid oxygen
mass transfer improvement of up to 6-fold (600%) at nanoparti-
cle volume fractions below 1% in an agitated, sparged reactor and
reported enhancement of both mass transfer coefcient (K
d
) and
the gasliquid interfacial area (a) in the presence of nanoparticles.
Interestingly, the enhancement in K
d
leveled off at a nanoparticle
volume fraction of about 1% (v/v). In another study [15], however,
contrary to the above, gasliquid mass transfer coefcient in a
three-phase internal loop airlift reactor was reduced in the pres-
ence of TiO
2
nanoparticles. They blamed increased aggregation of
nanoparticles due to increased particle holdup, to be responsible
for adverse mass transport.
As far as hydrodynamic characteristics are concerned, Kimet al.
[16] examined the bubble behaviors during the bubble absorp-
tion process for 8.0% ammonia solution. They reported spherical
but smaller bubble sizes, increased residence time and 22% lower
speed of bubble rise in the presence of Cu nanoparticles of 0.1wt%,
while the bubble shape was hemi-spherical in the absence of
nanoparticles. Fan et al. [17] studied the anomalous gas holdup
and bubble behavior in a bubble column or a micro-channel with
the presence of silicon dioxide (SiO
2
) nanoparticles in water. They
observed reduction in bubble sizes and increase in the gas holdup
in the bubble column in the presence of nanoparticles at super-
cial gas velocities above 710cm/s. In another study, Feng et al.
[18] investigated volume fraction of 1.13.3% hydrophobic TiO
2
nanoparticles inwater andcontrary to the above reporteda general
trend of decreased gas holdup with increasing volume fraction of
the nanoparticles.
As can be seen the above work on the application of nanouids
onmass transfer has dealt withgasliquidsystems, wherenanopar-
ticles are added to the liquid phase. Since hydrodynamics play an
important role on mass transfer operations, and considering the
mainadvantages of pulsed liquidliquid extractioncolumn(PLLEC)
[19,20] it was decidedtousethis columninthis studywhichis capa-
ble of offering a more roust mediumof investigations for studying
the inuence of nanoparticles on mass transfer and hydrodynamic
characteristics.
Three forms of holdup have been discussed in the literature:
static, dynamic, and total holdup [21]. static holdup refers to that
portion of the dispersed phase which is being trapped under the
discs [22] while dynamic holdup refers to the moving fraction of
the dispersed phase; these are being obtained by shut-down pro-
cedure; total holdup on the other hand, equals the sumof both.
2. Experimental
2.1. Experimental setup
Schematic diagramof the PLLEC used in this study is presented
in Fig. 1. The specication of PLLEC and range of operating variables
Please cite this article in press as: A. Bahmanyar, et al., The inuence of nanoparticles on hydrodynamic characteristics and mass transfer
performance in a pulsed liquidliquid extraction column. Chem. Eng. Process. (2011), doi:10.1016/j.cep.2011.08.008
ARTICLE IN PRESS
GModel
CEP-6047; No. of Pages 9
A. Bahmanyar et al. / Chemical Engineering and Processing xxx (2011) xxxxxx 3
Table 1
The specication of PLLEC and range of operating variables.
Column material Pyrex
Column height (cm) 70
Internal diameter of the column (mm) 90
Plate material Stainless steel 316
Plate diameter (mm) 88
Plate thickness (mm) 1
Number of sieve plates 10
Number of holes in plate 170
Plate hole diameter (mm) 3
Distance between hole pate (mm) 6
Hole pitch (mm) 6
Average free area of the plates (%) 20
Compartment height (cm) 5
Mass ux ratio (Qc/Q
d
) 1.2
Compressed air pressure (KPa) 60300
Range of pulsation intensity (cm/s) 0.32.3
are listed in Table 1. The column consists of a 70cm long vertical
Pyrex tube, i.d. =90mm, 10 perforated stainless steel plates reg-
ularly spaced by 5cm apart and supported on a 8mm diameter
central rod, as well as two separating chambers at either sides of
the column. Holes of 3mmdiameter were arranged in a 6mmtri-
angular pitch at each plate, providing a nominal free fraction area
of 0.20. Pulsation was obtained via a newly designed pulsator, con-
sisting an air compressor, a 3way-2 position-direct acting solenoid
valve and a micro controller (AVR-8051) as a programmable con-
troller unit (PCU) capable of energizing periodically the solenoid
valve to allow ow of the compressed air for an adjustable period
of time, thus pulsating the owalong the column with an intensity
of 0.32.3cm/s. The PCU provided adjustable time-off in the range
of 0.02065.535s. This pulsator which was specically designed
and built for this type of research works offers the advantages of
1ms accuracy in pulsation time, and does not require maintenance
of mechanical pump pulsators, nor suffers from the limitations of
airlift pulsators. Two ow meters were employed to supply and
monitor the xed ow rates of continuous and dispersed phases.
Photographs were taken using a Canon G9 digital camera (12.1
Mpix, ZoomLens 6 IS, 7.444.4mm, 1:2.84.8).
2.2. Experiments procedure
Both continuous and dispersed phases were mutually saturated
before being used in the experiments. At the beginning of each test
run, the continuous phase was initially allowed into the column
from the top which was lled up to the specied height (valve 4,
H=44.5cm). Dispersed phase was then fed into the column via a
glassy nozzle (inside diameter 2mm) from the bottom. The ow
Fig. 2. UVvis spectrumof SiO
2
nanoparticles in kerosene suspension.
Table 2
Description of chemical systems used.
Systems name Continuous phase Dispersed phase
WK SW
a
SK
WAAK SW SK+AA
WNF
1
SW SK+AA+0.01vol% HDK H18
WNF
2
SW SK+AA+0.05vol% HDK H18
WNF
3
SW SK+AA+0.1vol% HDK H18
WNF
4
SW SK+AA+0.01vol% HDK H20
WNF5 SW SK+AA+0.05vol% HDK H20
WNF
6
SW SK+AA+0.1vol% HDK H20
a
SW: saturated water; SK: saturated kerosene; AA: 0.05vol% acetic acid; NF:
nanouid.
meters are then xed to the specied amounts (Q
c
/Q
d
=1.2). Pulsa-
tion amplitude and frequency were then adjusted to desired values
by setting the time-on and time-off periods in PCU.
To investigate droplets behavior, the column was divided into
four different sections of 14 starting fromspaces betweenthe bot-
tom of the column and 1st valve, 1st and 2nd valves, 2nd and 3rd
valves, and 3rd and 4th valves, respectively. Apart from section 1
which had only one stage, each section consisted of three stages.
Once achieving steady state, photographs were taken of the
droplets using continuous shooting method at different sections
along the column, using the digital camera. At the end of a run,
the pulsation was turned off and the dynamic holdup was deter-
mined by the shut down method. In this method, after the system
was stabilized to allow steady state to be reached. Then, the inlet
and outlet valves were shut simultaneously and the moving frac-
tion of dispersed phase was allowed to coalesce at the interface.
The dynamic holdup was then measured either by determining
the change of interface height or displacing the solvent layer into
graduated cylinder. Finally, strong pulsation (2.3cm/s) was used to
move that portion of the dispersed phase which was being trapped
under the discs, and the static holdup was then measured.
2.3. Chemical systems
Acetic acid in the order of 5vol% was added to the saturated
kerosenetoprovidethebaseuidinall systems under investigation
listed in Table 2, except for the WK system. The nanouid was
prepared by dispersing two kinds of SiO
2
nanoparticles supplied by
WackerChemie Company (HDK H20 and HDK H18 with different
hydrophobicities of 20 and 18, respectively.) into the base uid.
To ensure efcient stability of nanoparticles, Hielscher ultrasound
generator (24kHz, 400W) was used for 1h duration (using H14
sonotrode with125m, 105W/cm
2
and 0.7s pulse duration).
Fig. 3. Linear relationship between light absorption and concentration of H18
nanoparticles in kerosene suspension at wavelength of 356.8nm.
Please cite this article in press as: A. Bahmanyar, et al., The inuence of nanoparticles on hydrodynamic characteristics and mass transfer
performance in a pulsed liquidliquid extraction column. Chem. Eng. Process. (2011), doi:10.1016/j.cep.2011.08.008
ARTICLE IN PRESS
GModel
CEP-6047; No. of Pages 9
4 A. Bahmanyar et al. / Chemical Engineering and Processing xxx (2011) xxxxxx
Fig. 4. Relative supernatant particle concentration of nanouids with respect to the
sediment time.
The amorphous fumed hydrophobic and highly hydrophobic
silica powders (HDK H18 and HDK H20, respectively) are highly
pure with nanoparticles in the range of 530nm and density of
2200kg/m
3
. Their surfaces have been modied by OSi(CH
3
)
2

groups. Densities of water and kerosene are 996 and 800kg/m


3
,
respectively. In addition, viscosities of the mentioned liquids are
1.0 and 1.67mPa s, respectively.
3. Results and discussion
3.1. Nanouids stability
The stability of nanouid was investigated with respect to
the sediment time using a UVvis spectrophotometer (UVICAM
8700 Series). The peak absorbance of both nanouids for WNF
3
and WNF
2
systems in saturated kerosene-based suspensions
presented in Fig. 2 appeared at 356.8nm. The absorbance of
nanoparticles decreased with increasing sediment time. A linear
relation as illustrated in Fig. 3 was obtained between the super-
Fig. 6. Sphericity of droplets versus volume fraction of nanoparticles at the 3rd
stage; PI =1.3cm/s, mass ux ratio=1.2.
natant concentration and the absorbance of suspended particles
fromwhich, the relative stability of nanouids was estimated with
respect to the sediment time.
Fig. 4 depicts the colloidal stability of the nanouids where after
8h, relative concentrations were maintained over 96% compared
with the initial concentrations, indicating acceptable stability of
the nanouids used in this work.
3.2. Inuence of nanoparticles on the droplet geometry
Experimental observations indicated that addition of various
contents of nanoparticles had a marked inuence on the geometri-
cal shapes of the droplets. To illustrate this, Fig. 5 demonstrates the
droplets formed at the 3rd stage, in similar operating conditions
of PI =1.3cm/s, mass ow ratio 1.2, with different nanoparticles
contents of 0.0, 0.01, 0.05 and 0.1vol% corresponding to WAAK,
WNF
1
, WNF
2
and WNF
3
, respectively.
As can be seen, droplets formed at zero or low nanoparti-
cles content (WAAK or WNF
1
) are ellipsoidal and those at
Fig. 5. Illustration of the inuence of SiO
2
nanoparticles content on the geometrical shape of droplets in the PLLEC (WAAK: 0.0vol%, WNF
1
: 0.01vol%, WNF
2
: 0.05vol%,
WNF
3
: 0.1vol%).
Please cite this article in press as: A. Bahmanyar, et al., The inuence of nanoparticles on hydrodynamic characteristics and mass transfer
performance in a pulsed liquidliquid extraction column. Chem. Eng. Process. (2011), doi:10.1016/j.cep.2011.08.008
ARTICLE IN PRESS
GModel
CEP-6047; No. of Pages 9
A. Bahmanyar et al. / Chemical Engineering and Processing xxx (2011) xxxxxx 5
Fig. 7. Dispersed phase static hold-up versus the pulsation intensity for different
chemical systems; mass ux ratio=1.2.
WNF
3
with 0.1vol% nanoparticles is almost spherical in shape,
indicating a trend in changing shape from ellipsoidal to spherical
droplets as nanoparticles content increased. Similar behavior has
been reported [16] for bubble columns in the presence of nanopar-
ticles. On the other hand, droplet sphericity (dened as the ratio of
the length of minor axis over the major axis), corresponds to higher
interfacial tension. This indicates that addition of nanoparticles
may havecontributedtothechangingshapetrendexplainedabove.
Very good explanation can be found for increasing the sphericity of
a droplet with interfacial tension [23].
Sphericity was measured for some 200 droplets at each
nanoparticles vol% (from0.00 to 0.1) and an average was taken as
demonstrated in Fig. 6, where almost complete spherical droplets
was observed at 0.1vol%.
3.3. Dispersed phase static hold-up
The experiments were designed for the PLLEC to operate at
no ooding conditions and in accordance to prescribed operating
condition recommended in the literature [24]. In the analysis of
data obtained froma PLLEC, conventionally pulsation intensity (PI),
dened as product of pulsation frequency, f, and stroke length or
pulsation amplitude, A, with the unit of (cm/s) is used.
PI = A f (1)
Fig. 7 shows the static hold-up versus the pulsation intensity
for different chemical systems used in this study for the given
mass owratio of 1.2. As expected, for all chemical systems under
investigation, maximum static hold-up was observed at the low-
est PI (namely 0.3cm/s). Maximum hold-up (7.08% at PI =0.3)
and minimumhold-up (0.22% at PI =2.3) observed throughout the
experiments, occurred for WNF
3
and WAAK systems, respec-
tively (see Fig. 8).
It should be noted that apart from the WK system, the static
hold-upreductiontookplacemorerapidlybyincreasingthePI upto
1.3. As canbe seenfromFig. 7, generally, static hold-updecreasedto
an asymptotic minimumwith increasing PI. This could well be due
to breakage of dispersed phase trapped droplets passing through
plate holes, and also conrms that the experimental set-up has
operated correctly. Different behavior of the WK system is most
likely due to lack of interactive inuence of both nanoparticles and
acetic acid in this system.
As far as the inuence of the nanoparticles presence in the sys-
tems studied is concerned, it can be seen that there is a marked
difference in the static hold-up of studied systems with different
nanoparticles contents subjected to various PIs. This difference in
static hold-up is more apparent for lower PIs, indicating more sig-
Fig. 8. Maximumand minimumhold-ups observed throughout the experiments for (a) WAAK and (b) WNF
3
.
Please cite this article in press as: A. Bahmanyar, et al., The inuence of nanoparticles on hydrodynamic characteristics and mass transfer
performance in a pulsed liquidliquid extraction column. Chem. Eng. Process. (2011), doi:10.1016/j.cep.2011.08.008
ARTICLE IN PRESS
GModel
CEP-6047; No. of Pages 9
6 A. Bahmanyar et al. / Chemical Engineering and Processing xxx (2011) xxxxxx
Fig. 9. Dispersedphasestatic hold-upfor all systems at PIs of (a) 1.3and(b) 1.8cm/s;
mass ux ratio=1.2.
nicant effect inthis pulsationrange due tonanoparticles presence.
In order to compare more clearly the impact of nanoparticles vol-
ume fraction and investigate the effect of hydrophobicity, the plot
of static hold-ups for all systems at PIs of 1.3 and 1.8cm/s were
prepared as presented in Fig. 9.
As for hydrophobicity, increase in it caused static hold-up to
increase for both PIs. Also considering Figs. 7 and 9 it may be
claimed that, higher volume fraction of nanoparticles led to an
increase in static hold-up for PI =1.3 or less, while it made little
difference when PI =1.8 or higher.
Increase in static hold-up caused by higher nanoparticles con-
tent, could be attributed to increased coalescence of droplets under
the plates. The coalescence of drops was observed under the plates
Fig. 11. Dispersed phase dynamic hold-up versus the pulsation intensity for differ-
ent chemical systems; mass ux ratio=1.2 (d
32
is expressed for WNF
3
).
in this experiment. It is mentioned by previous researchers that
the larger the interfacial tension, the more easily coalescence of
drops will occur [25]. So, Improved interfacial tension (as discussed
in Section 3.2) might be responsible for coalescence tendency of
droplets under the plates.
Fig. 10 illustrates the extent of this coalescence taken place
at Plate 3 in the presence of nanoparticles and compares it with
WAAK where no nanoparticles existed in the system.
As mentioned before, the increase in static hold-up with
nanoparticles content was more apparent for lower PIs (as con-
rmed by Fig. 7). This might be explained by the fact that high
PI provides such intensive turbulence and consequently more
droplets breakage that it supersedes the impact of coalescence
caused by the presence of nanoparticles.
3.4. Dispersed phase dynamic hold-up
Fig. 11 illustrates the dynamic hold-up versus PI for different
chemical systems at the mass owratio of 1.2. As shown, dynamic
hold-up generally increased with increased PI. Based on experi-
mental observation (not explained here for the sake of brevity)
obtainedfromthe photographic technique, the meandroplet diam-
eter decreased with higher PI leading to longer droplets resident
time. In other words, enhanced dispersed phase dynamic hold-up
was caused by higher population density of droplets inside the
column imposed by increased PI. This phenomenon continued up
to the highest hold-up attempted in the experiment and corre-
sponded to the minimumdrop size (see for example d
32
for WNF
3
expressedinFig. 11) obtainedunder available operatingconditions.
Fig. 10. Extent of coalescence taken place at Plate 3 for (a) WAAK (0.0vol%) and (b) WNF
3
(0.1vol%) systems.
Please cite this article in press as: A. Bahmanyar, et al., The inuence of nanoparticles on hydrodynamic characteristics and mass transfer
performance in a pulsed liquidliquid extraction column. Chem. Eng. Process. (2011), doi:10.1016/j.cep.2011.08.008
ARTICLE IN PRESS
GModel
CEP-6047; No. of Pages 9
A. Bahmanyar et al. / Chemical Engineering and Processing xxx (2011) xxxxxx 7
Fig. 12. Dispersed phase static hold-up for all systems at PI =1.3; mass ux
ratio=1.2.
As can be seen from Fig. 11, dynamic hold-up increased for
higher nanoparticles content. There are two possible explanations
for this which are associated with the increased density of the dis-
persedphase andenhancedopportunity for droplet breakage. With
regards to the former, increase in nanoparticles content, causes
higher density of the dispersed phase, which leads to increased
residence time and consequently increased dynamic hold-up. To
interpret the latter, as previously expressed, nanoparticles content
increased static hold-up as very large droplets coalesced in their
presence, thereby covering almost the entire plate surface in a sin-
gle layer (see Fig. 10). Subsequent pulsations therefore, have more
opportunity to break up these droplets into smaller ones leading
to increased dynamic hold-up. On the other hand, in the absence
of nanoparticles, (or lowquantity say 0.01vol%), the droplets were
mostly in the formof small scattered ones attached under the plate
with much less coverage of the plate holes, consequently, giving
less opportunity for droplet breakage in the following pulsations.
To compare more clearly the impact of nanoparticles volume
fraction, and investigate the effect of hydrophobicity, the plot of
dynamic hold-up for all systems at PI of 1.3 was prepared as
presented in Fig. 12. As for hydrophobicity, increase in it made
little increase in the dynamic hold-up while similar trend to H18
was observed for H20 nanoparticles, too. Similar results were also
obtained when PI changed from1.3 to 1.8.
3.5. Mass transfer performance
3.5.1. Theoretical background in calculation of experimental
dispersed phase mass transfer coefcients
Theoverall mass transfer of dispersedor continuous phaseis one
of the fundamental parameters in liquidliquid extractor design. In
this study, mass transfer coefcients were calculated using a semi-
empirical method described below. The mass balance for droplet
may be written as:
K
d
(C C

) 4r
2
=
4
3
r
3
dc
dt
(2)
Rearranging and integrating gives:
K
d
=

d
6t

ln(1 E) (3)
where
E =
C
0
C
C
0
C

(4)
C
0
, C, and C* are the solute concentration in the primary droplet
(before contact), at a specied height from the bottom of the col-
umn (44.5cm for the set-up in this work taken from valve 4), and
Fig. 13. Effect of PI onenhancement factor for different chemical systems; mass ux
ratio =1.2.
in equilibriumwith the continuous phase, respectively. The solute
concentrations of collected droplets were determined by titration
with 0.1N NaOH solution.
The relationbetweencontact time (t), dispersedphase volumet-
ric owrate (Q
d
), and hold-up () is as follows [25]:
L =
Q
d
t
S
(5)
When the experimental data (, d
32
, and Q
d
) throughout the
height of columnareavailable, as is thecaseinthis work, thecontact
timebetweentwo phases wereobtainedusingEq. (5). Mass transfer
coefcient (K
d
) along the column height could then be obtained
using Eqs. (3) and (4).
3.5.2. Determination of mass transfer enhancement and
interfacial area
Fig. 13 shows the effect of PI on the enhancement factor (E).
As demonstrated, mass transfer in the presence of the nanouids
has been improved by 421%, 547% and 560% for 0.01, 0.05 and
0.1vol%, respectively. This enhancement could be attributed to the
Brownian motion of nanoparticles. The gure also shows that for
WAAKsystem, Esignicantlyimproves whenPI is increasedfrom
1.3 to 1.8cm/s. The same trend is observed for WNF
1
systemhow-
ever, enhancement in E levels off at the latter PI. This leveling off
could be contributed to the substantial decrease of mass transfer
rates due to small droplets, behaving as rigid spheres, in which
case molecular diffusion would be governing mass transfer in the
system.
Moreover, a clear peak for both WNF
2
and WNF
3
systems are
observed at 1.8 and 1.3cm/s, respectively, suggesting that in the
presence of nanouids, mass transfer rate may not improve above
a certain PI (1.8cm/s in the present experiment). Nanoparticles,
on the other hand may have facilitated increasing the regidica-
tion of small drops which may have caused the peak for WNF
3
(0.1vol%) to occur at lower pulsation intensity compared to WNF
2
(0.05vol%).
Theinterfacial areaavailableinacounter-current extractioncol-
umn depends upon the volume fraction or dynamic holdup, of the
dispersed phase, as well as on the mean droplet size and can be
evaluated by [24]:
a =
6
d
d
32
(6)
As it is explained in Section 2.2 in each experiment, the sauter
mean diameter. d
32
, was obtained by photographic technique at a
Please cite this article in press as: A. Bahmanyar, et al., The inuence of nanoparticles on hydrodynamic characteristics and mass transfer
performance in a pulsed liquidliquid extraction column. Chem. Eng. Process. (2011), doi:10.1016/j.cep.2011.08.008
ARTICLE IN PRESS
GModel
CEP-6047; No. of Pages 9
8 A. Bahmanyar et al. / Chemical Engineering and Processing xxx (2011) xxxxxx
Fig. 14. Effect of PI on interfacial area for different chemical systems; mass ux
ratio=1.2.
few stages of column height and then calculated by following Eq.
(7):
d
32
=

n
i
d
3
i

n
i
d
2
i
(7)
When the experimental data (, d
32
) are available, as is the
case in this work, the interfacial was obtained using Eq. (6). The
effect of PI on the interfacial area is illustrated in Fig. 14, where
interfacial area increased with an increase in PI. Dynamic holdup
and Sauter mean diameter are function of physical properties and
operating conditions. So, due to above mentioned equations dif-
ferent variation trends could be expected for different systems in
Fig. 14. As mentioned previously, at higher PI, sauter mean diam-
eter decreased and consequently dynamic hold-up increased (see
Fig. 11), both of which may have contributed to the increased inter-
facial area observed.
Table 3 shows the mass transfer coefcient (K
d
) for vari-
ous systems tested, where evidently K
d
generally decreased with
increasing PI. Experimental observations, indicated that increas-
ing PI, decreased d
32
and consequently increased resident time (t).
However, as expressed by Eq. (2), K
d
decreases with both effects.
Having established experimentally that nanoparticles presence
do affect parameters such as enhancement factors, static and
dynamic hold-ups, it was expected to observe more clearly their
inuence on K
d
, which was not the case here. Considering con-
icting reports from the available literature on the inuence of
nanoparticles on K
d
[9,14,15], and mass transfer performance, it
could only be said that the inuence is less clear on K
d
as compared
to the mass transfer performance. This suggest the need for further
research to be carried out into modeling exercise of mass trans-
fer coefcient in the presence of nanoparticles, for liquidliquid
systems, similar to those carried out by Nagy et al. [26] on the
gasliquid systems.
Table 3
Experimental value of mass transfer coefcient at different PI.
Pulse intensity
PI (cm/s)
Mass transfer coefcient for different systems,
k
d
10
5
(m/s)
WAAK
a
WNF
1
a
WNF
2
a
WNF
3
a
0.3 7.10 3.11 4.21 4.44
0.7 3.15 2.56 2.72 2.55
1.3 2.71 1.94 2.01 1.87
1.8 1.48 1.26 1.28 0.90
2.3 1.06 0.69 0.81 0.62
a
Systems name, WAAK, WNF
1
, WNF
2
, WNF
3
, are dened in Table 2.
4. Conclusion
To investigate the inuence of nanoparticles on the hydrody-
namic characteristics and mass transfer performance of pulsed
liquidliquid extraction column different nanouids have been
applied as dispersed phase. From the experimental results, it can
be seen that there is a marked difference in the static and dynamic
hold-ups as well as enhancement factor of studiedsystems withdif-
ferent nanoparticles contents subjected to various PIs. It is found
that static and dynamic dispersed phase hold-ups for the nanou-
ids increased 23398%, and 23257%, respectively while the mass
transfer for the nanouids enhances 460%. The increase in static
hold-up and enhancement factor with nanoparticles content was
more apparent for lower PIs indicating more signicant effect
in this pulsation range due to nanoparticles presence. As far as
hydrophobicity is concerned, little increases in both static and
dynamic holdups were observed as hydrophobicity increased and
interestingly, similar trend was observed for both H18 and H20
nanoparticles.
Appendix A. Nomenclature
a interfacial area (m
2
/m
3
)
A pulsation amplitude (cm)
C solute concentration in dispersed phase (kg/m
3
)
C
0
initial concentration of solute in dispersed phase (kg/m
3
)
C* equilibrium concentration of solute in dispersed phase
(kg/m
3
)
d droplet diameter (m)
d
32
sauter mean diameter of droplets (mm)
E enhancement factor
f pulsation frequency (s
1
)
H effective height of column (cm)
K
d
dispersed phase mass transfer coefcient (m/s)
L column height (m)
PI pulsation intensity (cm/s)
Q
c
continuous phase owrate (m
3
/s)
Q
d
dispersed phase owrate (m
3
/s)
R
1
length of major axis (mm)
R
2
length of minor axis (mm)
S column cross-sectional area (m
2
)
t resident time of dispersed phase in column, contact time
(s)
Greek symbols
dispersed phase hold-up
Subscripts
c continuous phase
d dispersed phase
References
[1] J.A. Eastman, S.U.S. Choi, S. Li, W. Yu, L.J. Thompson, Anomalously increased
effective thermal conductivities of ethylene glycol based nanouids containing
copper nanoparticles, Appl. Phys. Lett. 78 (2001) 718720.
[2] S.K. Das, N. Putra, P. Thiesen, W. Roetzel, Temperature dependence of thermal
conductivity enhancement for nanouids, Trans. ASME, J. Heat Transfer 125
(2003) 567574.
[3] S.Z. Heris, S.G. Etemad, M.N. Esfahany, Experimental investigation of oxide
nanouids laminar ow convective heat transfer, Int. Commun. Heat Mass
Transfer 33 (2006) 529535.
[4] W. Yu, D.M. France, J.L. Routbort, S.U.S. Choi, Review and comparison of
nanouid thermal conductivity and heat transfer enhancements, Heat Trans.
Eng. 29 (2008) 432460.
[5] Y.M. Xuan, Q. Li, Investigation on convective heat transfer and owfeatures of
nanouids, Trans. ASME, J. Heat Transfer 125 (2003) 151155.
Please cite this article in press as: A. Bahmanyar, et al., The inuence of nanoparticles on hydrodynamic characteristics and mass transfer
performance in a pulsed liquidliquid extraction column. Chem. Eng. Process. (2011), doi:10.1016/j.cep.2011.08.008
ARTICLE IN PRESS
GModel
CEP-6047; No. of Pages 9
A. Bahmanyar et al. / Chemical Engineering and Processing xxx (2011) xxxxxx 9
[6] S.K. Das, S.U.S. Choi, H.E. Patel, Heat transfer innanouids a review, Heat Trans.
Eng. 27 (2006) 319.
[7] R. Prasher, P. Bhattacharya, P.E. Phelan, Brownianmotion-based convective-
conductive model for the effective thermal conductivity of nanouids, Trans.
ASME, J. Heat Transfer 128 (2006) 588595.
[8] S.P. Jang, S.U.S. Choi, Role of Brownian motion in the enhanced thermal con-
ductivity of nanouids, Appl. Phys. Lett. 84 (2004) 43164318.
[9] S. Krishnamurthy, P. Bhattacharya, P.E. Phelan, R.S. Prasher, Enhanced mass
transport in nanouids, Nano Lett. 6 (2006) 419423.
[10] H. Li, A. Prakash, A. Margaritis, M.A. Bergougnou, Effects of micron-sized par-
ticles on hydrodynamics and local heat transfer in a slurry bubble column,
Powder Technol. 133 (2003) 171184.
[11] G. Quicker, A. Schumpe, W.D. Dechwer, Gasliquid interfacial-areas in a bubble
column with suspended-solids, Chem. Eng. Sci. 39 (1984) 179183.
[12] M.V. Dagaonkar, H.J. Heeres, A.A.C.M. Beenackers, V.G. Pangarkar, The applica-
tion of ne TiO
2
particles for enhanced gas absorption, Chem. Eng. J. 92 (2003)
151159.
[13] Y.T. Kang, H.J. Kim, K.I. Lee, Heat and mass transfer enhancement of binary
nanouids for H
2
O/LiBr falling lmabsorption process, Int. J. Refriger. 3 (2008)
850856.
[14] B. Olle, S. Bucak, T.C. Holmes, L. Bromberg, A. Hatton, D.I.C. Wang, Enhancement
of oxygen mass transfer using functionalized magnetic nanoparticles, Ind. Eng.
Chem. Res. 45 (2006) 43554363.
[15] J. Wen, X. Jia, W. Feng, Hydrodynamic and mass transfer of gasliquidsolid
three-phase internal loop airlift reactors with nanometer solid particles, Chem.
Eng. Technol. 28 (2005) 5360.
[16] J.K. Kim, J.Y. Jung, Y.T. Kang, The effect of nano-particles on the bubble absorp-
tion performance in a binary nanouid, Int. J. Refriger. 29 (2006) 2229.
[17] L.S. Fan, O. Hemminger, Z. Yu, F. Wang, Bubbles in nanouids, Ind. Eng. Chem.
Res. 46 (2007) 43414346.
[18] W. Feng, J. Wen, J. Fan, Q. Yuan, X. Jia, Y. Sun, Local hydrodynamics of gasliquid-
nanoparticles three-phase uidization, Chem. Eng. Sci. 60 (2005) 68876898.
[19] K. Gottliebsen, B. Grinbaum, D. Chen, G.W. Stevens, The use of pulsedperforated
plate extraction column for recovery of sulphuric acid fromcopper tank house
electrolyte bleeds, Hydrometallurgy 58 (2000) 203213.
[20] Y. Wang, S. Jing, G. Wu, W. Wu, Axial mixing and mass transfer characteristics
of pulsed extraction Column with discs and doughnuts, Trans. Nonferrous Met.
Soc. Chin. 16 (2006) 178184.
[21] S.S. Puranik, A. Vogelpohl, Effective interfacial area inirrigatedpackedcolumns,
Chem. Eng. Sci. 29 (1974) 501507.
[22] M.J. Brodkorb, M.J. Slater, Multicomponent and contamination effects on mass
transfer in a liquidliquid extraction rotating disc contactor, Trans. IChemE 79
(2001) 335345.
[23] D. Mobius, R. Miller, Drops andBubbles inInterfacial Research, Elsevier Science,
NewYork, 1998, pp. 189197.
[24] J.C. Godfrey, M.J. Slater, LiquidLiquid Extraction Equipment, 2nd ed., John
Wiley and Sons, NewYork, 1994, pp. 227305.
[25] R.E. Treybal, Mass Transfer Operations, 3rd ed., McGraw Hill, Japan, 1990, pp.
488490.
[26] E. Nagy, T. Feczk, B. Koroknai, Enhancement of oxygen mass transfer
rate in the presence of nanosized particles, Chem. Eng. Sci. 62 (2007)
73917398.

Você também pode gostar