Bubble Column Reactors For Wastewater Treatment. 3. Pilot-Scale Solvent Sublation of Pyrene and Pentachlorophenol From Simulated Wastewater

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Ind. Eng. Chem. Res.

1997, 36, 903-914 903

Bubble Column Reactors for Wastewater Treatment. 3. Pilot-Scale


Solvent Sublation of Pyrene and Pentachlorophenol from Simulated
Wastewater
Jeffrey S. Smith† and Kalliat T. Valsaraj*
Department of Chemical Engineering, Louisiana State University, Baton Rouge, Louisiana 70803

Removal efficiencies from a three-phase continuous, countercurrent solvent sublation of pyrene


and pentachlorophenol from water are presented and compare well with Series CSTR Model
predictions. The effects of the air-to-water flow ratio, the solvent flow rate, and the gas sparger
design are discussed. Steady-state efficiencies as high as 96% for pyrene and 94% for
pentachlorophenol are reported for an annular shear gas sparger. These results demonstrate
the superiority of solvent sublation over bubble fractionation which is shown to be limited to a
removal efficiency of 68% or less for these compounds. Moreover, analyses of the solvent (mineral
oil) phase show that the separation factor (ratio of the solvent effluent concentration to the
water effluent concentration) in solvent sublation can be as high as 300, which is distinctly
greater than that observed in solvent extraction.

Introduction is that a nonvolatile organic solvent is floated upon the


aqueous phase. The advantage of the solvent is that it
In wastewater treatment plants, aeration devices, serves as a sink for VOCs as well as nonvolatile organic
such as diffused aerators and mechanical surface aera- compounds (NVOCs) which are resistant to biodegrada-
tors, are often used to facilitate the biological oxidation tion. Moreover, in the sublation process, the surface-
of organic compounds. In the case of diffused aeration, active nature of organic compounds is exploited. This
the process closely resembles that of a short, squat phenomenon has a profound favorable effect on the
bubble column reactor with multiple air injection points. removal efficiencies of these compounds, in particular
Recently (Chern and Yu, 1995), it has been reported that strongly hydrophobic compounds. The reasons why
the air stripping which takes place in these devices is a efficiencies are higher in sublation than in other separa-
serious and growing concern as waste treatment opera- tion processes are because (i) the adsorbed phase and
tors are being charged to further control their volatile the bubble wake, which is entrained into the overlying
organic compound (VOC) emissions. The authors solvent, experience solvent extraction, (ii) the overall
stressed the need for improved mass-transfer models driving force for mass transfer in the air-water disper-
which describe VOC emission from these devices and sion is increased, and (iii) the mechanism of adsorption
introduced a new oxygen mass-transfer model. Al- to the bubble-water interface circumvents the air-side
though such activities are indeed very good first steps, mass-transfer resistance which is the dominant resis-
the problem of preventing VOC emission from these tance for hydrophobic compounds (of low vapor pres-
devices still remains. Moreover, the current system fails sures) in conventional air-water contactors.
to remove organic compounds which are resistant to In this paper, results obtained from a pilot-scale
biological oxidation such as polynuclear aromatic hy- sublation process used to treat aqueous streams con-
drocarbons and chlorinated pesticides and insecticides. taining pyrene and pentachlorophenol are reported. The
Though regulations may not require action now, it is effects of the air-to-water flow ratio, the gas sparger
anticipated that new or improved aeration technology design, and the Henry enhancement factor on fractional
that reduces VOC emission as well as targets com- removal are discussed. The Henry enhancement factor,
pounds which resist biodegradation will soon be re- which was defined in part 1, is the ratio of the effective
quired. Liquid-liquid extraction is an alternative for Henry law constant, which arises from the combination
the removal of nonvolatile compounds if an appropriate of bulk phase partitioning and surface adsorption, to the
solvent is available (Cusack, 1996). Solvent extraction, conventional Henry law constant. The results show
however, leaves significant residual solvent in the water that solvent sublation is distinctly superior to bubble
which requires a subsequent separation step. One fractionation (diffused aeration). Furthermore, the
potential alternative to conventional solvent extraction Series CSTR Model (SCM) developed in Part 1 is tested
is solvent sublation. for predicting sublation performance.
The solvent sublation process, which was introduced
in parts 1 and 2 of this series (Smith et al., 1996a,b), is Experimental Section
a wastewater treatment technology that has the poten-
tial to remove volatile and nonvolatile compounds Materials and Methods. The equipment used in
without the problem of residual solvent. Like the this investigation included a 4 in. i.d. × 60 in. long
diffused aeration device, solvent sublation closely re- (0.1016 × 1.524 m) sublation column, equipped with
sembles a squat bubble column reactor and is basically both a fine porous glass frit and an annular shear
operated in the same way; however, the only difference sparger (0.1 ft2, 6.3 ft/s water jet) described in part 2 of
this series. By using the two different spargers, we were
* To whom correspondence should be addressed. E-mail: able to investigate the effect of the different column
kvalsar@lsuvm.sncc.lsu.edu. Telephone: (504) 388-6522. hydrodynamics. Pyrene and pentachlorophenol (PCP)
† Separations Expertise Center, Dow Corning Corp. No. 128, were selected as the target organics to be removed from
Midland, MI 48686. E-mail: usdccmmj@ibmmail.com. the aqueous phase. Pyrene is an example of a poly-
S0888-5885(96)00524-6 CCC: $14.00 © 1997 American Chemical Society
904 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

Table 1. Operating Conditions and Steady-State Fractional Removal Results for the Solvent Sublation and Bubble
Fractionation of Pyrene and Pentachlorophenol
run I.D. compound sparger instr./control QA, SCCM Qw, cm3/min Qo, cm3/min FRss
Sublation
R-73-9-21-94 pyrene porous frit ultrasonic 959 41.8 4 0.809
R-79-9-27-94 pyrene porous frit ultrasonic 959 27.9 4 0.863
R-80-9-29-94 pyrene porous frit ultrasonic 959 41.8 4 0.796
R-87-10-12-94 pyrene porous frit ultrasonic 959 27.9 2 0.862
R-88-10-13-94 pyrene porous frit ultrasonic 2225 27.9 2 0.892
R-89-10-15-94 pyrene porous frit ultrasonic 4413 27.9 2 0.896
R-92-11-1-94 pyrene porous frit ultrasonic 443 27.9 2 0.771
R-94-11-15-94 pyrene porous frit ultrasonic 443 55.8 2 0.684
R-97-11-17-94 pyrene porous frit ultrasonic 1592 13.9 2 0.893
R-99-11-30-94 pyrene porous frit ultrasonic 2225 41.8 2 0.833
R-100-12-1-94 pyrene porous frit ultrasonic 8307 41.8 2 0.901
R-103-12-7-94 pyrene porous frit ultrasonic 959 27.9 2a 0.832
R-106-1-18-95 pyrene annular shear ultrasonic 2225 27.9 2 0.887
R-113-2-9-95 pyrene annular shear ultrasonic 4093 27.9 2 0.945
R-117-3-7-95 pyrene annular shear ultrasonic 5977 27.9 2 0.962
R-120-3-21-95 pyrene annular shear ultrasonic 5355 13.9 2 0.956
R-124-4-6-95 pyrene annular shear ultrasonic 2225 55.8 2 0.882
R-16-6-22-95 PCP (pH 9) annular shear ultrasonic 4732 27.9 2 0.127
R-24-9-20-95 PCP (pH 3) annular shear conductivity 4732 27.9 2 0.853
R-30-9-26-95 PCP (pH 3) annular shear conductivity 7156 27.9 2 0.907
R-33-9-29-95 PCP (pH 3) annular shear conductivity 7156 13.9 2 0.942
R-37-10-3-95 PCP (pH 3) annular shear conductivity 2840 55.8 2 0.744
R-40-10-9-95 PCP (pH 3) annular shear conductivity 2840 41.8 2 0.783
R-43-10-10-95 PCP (pH 3) annular shear conductivity 3454 27.9 2 0.858
Bubble Fractionation
R-131-6-13-95 pyrene annular shear ultrasonic 5977 27.9 no solvent 0.673
R-58-11-27-95 PCP (pH 3) annular shear conductivity 7156 27.9 no solvent 0.682
a R-103 was conducted with “spent solvent”; that is, the solvent used in R-100 was used again.

nuclear aromatic hydrocarbon (PAH) which is a byprod- Table 2. Settings for Chromatographic Equipment
uct from coal gasification, and PCP is an example of a Schimadzu Autoinjector
commercially manufactured insecticide. Both com- injection volume (µL) 25
pounds pose a significant threat to the environment and no. of replicates 3
are classified as priority pollutants by the U.S. Envi- HP 1090L HPLC (Reverse Phase)
ronmental Protection Agency under the Clean Water mobile phase 70-85% MeCN (variable)
Act of 1977. Light-white mineral oil was chosen as the stationary phase Phenomenex Envirosep-PP
solvent phase as it has several advantages over other (3.2 × 125 mm)
organic liquids. For instance, the vapor pressure is low; flow rate (mL/min) 0.5
the density is 0.86 g/cm3; and there is no foul odor. The pressure (bar) 100
retention time (min) 4.5
compounds used in study were obtained from Sigma-
Aldrich, while distilled water and air were supplied by HP 1046A Fluorescence Detector
excitation wavelength (nm) 237
LSU plant utilities. emission wavelength (nm) 385
Twenty-six experiments were conducted which inves- response time (ms) 5 (2000)
tigated the effects of the air, water, and solvent flow scan speed (nm/s) 6 (50)
rates, the gas sparger design, and the Henry enhance- photomultiplier gain 12
ment factor. A summary of the experiments and
operating conditions appears in Table 1. The water fed The detector was connected in series with a Hewlett-
to the process was prepared by spiking containers of Packard 3932A integrator, a Hewlett-Packard 1090L
distilled water (≈19 L each) with the target organic from high-performance liquid chromatograph, and a Shi-
gravimetrically prepared acetonitrile (MeCN) solutions. mazdu autoinjector. The settings for the analytical
For the pyrene experiments, the concentrations of the equipment are listed in Table 2. A three-step procedure
feed stocks ranged between 86 and 130 µg/L, while those was used for preparing aqueous standards for calibra-
for the PCP experiments ranged between 3 and 9 mg/ tion. First, a gravimetric solution of pyrene in MeCN
L, with most being equal to 6 mg/L. For those experi- (3192 mg/L) was prepared; second, 1000 µL of the
ments which investigated the removal of neutral PCP, gravimetric solution was diluted with 100 mL of MeCN
sulfuric acid was added to the feed stocks to adjust the to make a 30 mg/L solution; finally, the samples were
pH between 2.5 and 3.0. Since the pKA of PCP is 4.74 further diluted with distilled water to make solutions
(Montgomery and Welkom, 1990), this ensured that the ranging from 1 to 135 µg/L. The relative standard
PCP remained neutral throughout the experiments. In deviation (RSD) of the fluorescent detector was 13%,
a typical experiment, aqueous influent and effluent which after error propagation corresponds to a 2.5%
samples were taken intermittently and analyzed for the RSD in fractional removal.
organic compound being sublated. Each analysis in- For the analysis of pyrene in mineral oil, a Hewlett-
cluded two replicates. Fractional removal was calcu- Packard 8452 UV/vis spectrophotometer was used. A
lated as FR ) 1 - Aeff/Ainf, where Aeff/Ainf is the ratio of standard solution was prepared gravimetrically by
the average aqueous, effluent concentration to the dissolving 0.5 g of solid pyrene in 1 L of mineral oil. To
average aqueous, influent concentration. calibrate the detector, aliquots of the standard were
A Hewlett-Packard 1046A fluorescent detector was diluted with fresh oil to make solutions ranging from
used to determine the concentration of pyrene in water. 0.5 to 7.40 ppm (wt). The analytical wavelength used
Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997 905
ultrasonic switch and adjusting the height of the weir
in the overflow reservoir.
To eliminate the need for human intervention, a
second control scheme, named DAC (data acquisition
and control), was constructed. In the DAC system, the
control action was based upon the electrical-conductivity
profile in the overflow reservoir. As illustrated in
Figure 2, six pairs of nichrome electrodes, encased in
glass tubes, were immersed in the overflow reservoir
to different depths. The probes were arranged to
maximize the sensitivity near the interface. Each pair
of probes was subjected to a 5 kHz ac potential. (High
frequency was necessary as it prevented polarization of
the probes.) A multimeter measured the conductances
across the probes and transmitted the information to a
PC where the interface position was calculated. A
model-based control algorithm then determined the
output to the solenoid valve. One of the unique features
of the controller was that it simulated the valve action
as if it had full dynamic range. (Details of the DAC
system are given in the Appendix, which is available
upon request.) With the DAC system, the control of the
process was improved significantly as deviations about
steady state were within (0.012 fractional removal
points. Figure 1c shows a representative run where
conductivity-based control was used.
Deoxygenation Experiments for the Determina-
tion of KLav. Oxygen-based mass-transfer coefficients
were determined with the sublation column operating
in a semibatch mode (Qo ) 0, Qw ) 0). A YSI Model 55
oxygen meter was used to measure the dissolved oxygen
in the water. In a typical experiment, air, at a known
Figure 1. Examples of fractional removal profiles obtained with
(a and b) ultrasonic Level Control and (c) conductivity-based level flow rate, was sparged into the column until the meter
control. showed that the dissolved oxygen was near 8.6 mg/L,
the solubility at 22 °C, and elevation at 0. (Distilled
for measuring absorbence was 340 nm. The uncertainty water was used.) Once steady state had been main-
of the measurements (95% C.I.) was 6.48%. tained for several minutes, the feed gas was switched
For the analysis of PCP in water, the UV spectropho- from instrument air to nitrogen. The switch to nitrogen
tometer was used to measure absorbence at 214 nm. A occurred quickly and smoothly without any fluctuation
520 mg/L solution of PCP in MeCN was prepared in the gas flow rate. This was possible because the two
gravimetrically, and aliquots of this standard were regulators controlling the source of air and nitrogen
diluted with distilled water to make solutions ranging were fed into one common regulator which fed the
from 0.5 to 10 mg/L. Fifty microliters of sulfuric acid column. During the experiment, the drop in dissolved
was added to each standard so that the pH would be oxygen was recorded with time. Plotting the data as
similar to the actual aqueous samples from the subla- ln Co/C versus time, as eq 1 suggests, resulted in a slope
tion column. The RSD for the fractional removal of equal to KLav
PCP was 1.5%.
Control Aspects. As described in part 2 of the Co 3k
series, the control of the solvent-water interface posi- ln t ∼ KLavt (1)
C
)
a(1 - )
tion at the top of the column required the most attention
as the flows to the process were all well controlled with Results and Discussion
metering pumps and flow tubes (rotameter). An ultra-
sonic level switch, which measured the time for sound Effect of Operational Variables (QA, Qw, Qo) on
to traverse the two phases, and a solenoid valve were Fractional Removal. The most important variable
used in most of the experiments for controlling the affecting the sublation of organic compounds from water
interface position. This arrangement served reasonably is the air-to-water flow ratio, QA/Qw. This ratio not only
well. Figure 1a shows a typical fractional removal influences the extent of stripping but also contributes
profile obtained from the sublation process with the greatly to column hydrodynamic properties. In part 2,
ultrasonic-based controller. On occasion, however, pro- it was argued that the homogeneous flow regime is the
cess upsets did occur and were the result of the appropriate operating regime for the sublation process.
ultrasonic switch failing to detect deviations from set- Thus, the values of QA and Qw shown in Table 1 were
point. For example, two upsets occurred during run chosen to ensure homogeneous flow. In Figure 3,
R-88 as shown in Figure 1b. In both occurrences, the fractional removal results obtained from the sublation
solvent was displaced with water because the switch process are plotted versus QA/Qw. Each point represents
failed to detect that the interface position had increased. a 15-30 h steady-state average. The general trend is
This type of failure was more common than solvent that, as the flow ratio is increased, the fractional
infiltration because of the large Qw/Qo ratios used in the removal rapidly approaches an upper limit. This be-
experiments. The process upsets were avoided with havior is consistent with SCM (Series CSTR Model)
simple manual adjustments that included resetting the predictions presented in part 1, which showed a very
906 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

Figure 2. Diagram of conductivity probes and data acquisition and control (DAC) system used to control the solvent-water interface.

Given the strong influence of the air-to-water flow ratio


observed in Figure 3, this was not surprising. When
the water flow rate was returned to its original value
(run R-80), the fractional removal returned to 80%. A
second experiment, run R-87, was performed which
investigated the effect of reducing the solvent flow rate
from 4 to 2 cm3/min while holding the water flow rate
constant at 27.9 cm3/min. As shown in Table 1, for runs
R-79 and R-87, there was no effect of reducing the
solvent flow rate on the fractional removal. The reason
for this is that the mineral-oil-partition coefficient (Kow)
for pyrene is very large. In fact, since most hydrophobic
compounds possess very large partition coefficients, the
amount of solvent needed for sublation is only that
which ensures a steady state. This was shown in part
1. Thus, the solvent flow rate used in most of the
experiments was maintained at the lowest pump set-
Figure 3. Fractional removal of pyrene and neutral pentachlo- ting, 2 cm3/min.
rophenol (PCP) versus the air-to-water flow ratio, QA/Qw. Because of the large capacity of the solvent, one
steep gradient between 0 e QA/Qw e 50 and a plateau variation of the process is to recycle the solvent. This
at QA/Qw g 150. In fact, the data in the figure suggest is referred to as using “spent” solvent and is a way of
the optimum operating condition is in the range of 200 minimizing the overall solvent usage. Therefore, to
e QA/Qw e 300, with the exact value depending on the further show the weak dependence of the solvent, run
compound and sparger type. Defining the optimum R-103 was conducted using a feed solvent (oil) having a
range is based on the observation that the gains made pyrene concentration of 1.57 mg/L; however, the operat-
in fractional removal are very small when QA/Qw is ing conditions of R-103 were identical to those used in
increased beyond 300. Another reason is that ulti- R-87. The effect of the spent solvent was to reduce the
mately, as the gas velocity is increased, an unfavorable fractional removal by only 0.03. Since this difference
transition occurs from homogeneous flow to heteroge- was outside the statistical error limits ((0.02), the effect
neous flow. The turbulence associated with heteroge- was statistically significant but, nonetheless, very weak.
neous flow is undesirable, as it tends to disperse the Effect of Gas Sparger on Mass Transfer. As
solvent throughout the system. illustrated in Figure 3, sublation experiments involving
Sublation experiments were performed to examine the pyrene were conducted which investigated the effect of
relative effect of the solvent flow rate versus the water gas sparger type. For QA/Qw < 150, there was no
flow rate. First, runs R-73 and R-79, which appear at significant difference between the fractional removal
the top of Table 1, were conducted to demonstrate the data collected from the shear sparger and those collected
effect of a -33% step change in water flow rate while from the porous frit. However, for QA/Qw g 150, the
holding the solvent flow rate constant at 4 cm3/min. As fractional removal of pyrene was improved by as much
one can see from the table, the fractional removal of as 5%, suggesting that the benefit of the shear sparger
pyrene increased from approximately 80% to 86% in is only realized at large QA/Qw ratios. The reason for
response to decreasing Qw from 41.8 to 27.9 cm3/min. this behavior may be explained by considering the
Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997 907
unique hydrodynamic properties of the shear sparger. bubble diameter, ds, divided by the rise velocity, U
It was observed in part 2, that, for gas velocities greater (Taylor and Krishna, 1993). This is because the liquid
than about 1 cm/s, the processes of bubble coalescence in immediate contact with a gas bubble is replaced in a
and breakage were buffered by the combined actions of time equal to that for the mean bubble to rise one
the water jet and the milky-white plume of microbubbles diameter. However, it is implied that the resistance to
exiting the shear sparger. It was speculated that this mass transfer resides totally in the liquid phase, that
so-called “buffering effect” was responsible for the weak is, the phase which experiences the penetration. Thus,
bubble size dependence on gas velocity. As it turns out, eq 3 is valid for single-phase resistance only. Addition-
in the cases where the shear sparger performed better ally, it is implied that the bubbles rise in pure plug flow
than the frit, the gas velocities varied from 0.84 to 1.23 since the effects of backmixing or circulation are ig-
cm/s. For the low QA/Qw experiments where the re- nored. In order to generalize, the key questions which
moval efficiencies were about the same for both sparg- must be addressed are as follows: How does one
ers, the gas velocities were less than or equal to 0.46 combine gas-phase resistance into eq 3? How does one
cm/s. To investigate whether or not the improved determine the exposure time when the flow regime is
performance was just the result of operating at the different from plug flow?
higher gas velocities, run R-100 was conducted using Sherwood et al. (1975) derived that the total moles
the porous frit sparger and a gas velocity of 1.71 cm/s transferred during an exposure period for two resis-
(QA/Qw ) 199). The result of this experiment was no tances in series may be expressed as

{x ( ( )
different than those observed at 0.33 (QA/Qw ) 115) and
0.91 cm/s (QA/Qw ) 158) for the porous frit. Thus, one 4DABte DAB κi2te
may reasonably conclude that the shear sparger does ∫0t N dt )
e

π
+
κi
exp
DAB
×

( x ) )}( )
have an advantage over the frit for superficial gas
velocities g 1 cm/s and that the effect is not simply the te Cg
result of gas velocity. The advantage is attributed to erfc κi -1 Cl - (4)
improved hydrodynamic properties of the shear sparger DAB Hc
and how they affect the air-water mass-transfer coef-
ficient. The following discussion elucidates this rela- where κi is the resistance at the interface due to a
tionship. surfactant. In the derivation of eq 4, the authors
Consider eq 2, which is the classical two-resistance assumed that the resistance 1/κi may be added to the
model for the overall liquid-side mass-transfer coef- liquid-side resistance in a fashion similar to the way
ficient. For many organic compounds such as VOCs, gas-side resistance was added to eq 2. Thus, eq 4 may
be used as a starting point to generate a general two-
1 1 1 phase penetration model describing the overall liquid-
(2) side mass-transfer coefficient. In the spirit of the
k κl Hcκg
) +
conventional film theory, we assume that the interface
itself offers no resistance but that the gas phase offers
air-side resistance is unimportant and the term 1/Hcκg
significant resistance. Substituting 1/Hcκg for 1/κi and
is neglected. However, for strongly hydrophobic com-
after some rearrangement, one can show
pounds such as pyrene and pentachlorophenol, air-side
resistance is important, if not the dominant resistance,

x
because of the very low Henry law constants, Hc, that 4DAB
they possess. Thus, the common practice of scaling k πte exp(z2) erfc(z) - 1
oxygen mass-transfer coefficients (which are liquid-side- (5)
Hcκg Hcκg
) +
controlled) by the Schmidt number is not correct for z2
these compounds. Therefore, other means must be used
to determine k. where z is defined as Hcκg(te/DAB)0.5, and Hc is the bulk
One common approach for predicting the mass- phase Henry law constant. When z is approximately
transfer coefficient for gas-liquid dispersions is to use greater than or equal to 1, the “correction function”, f(z),
a penetration theory such as the Higbie penetration becomes effectively zero and eq 5 reduces to eq 3;
model (Taylor and Krishna, 1993). In this model the however, if z is <1, then the effect of air-side resistance
average mass-transfer coefficient is defined as is significant.
The argument, z, depends upon the characteristic
length and velocity. It is proposed that the character-

x x
4DAB 4DAB
κl ) ) (3) istic length describing the exposure time of a packet of
πte λ fluid (eddy) in a bubble column is the Kolmogoroff
π
ω mixing length. Baird and Rice (1975) and Kantak et
al. (1994) have suggested expressions describing the
where DAB is the binary diffusivity and te is the exposure degree of mixing, or axial dispersion, in bubble columns
time. The exposure time may be considered as the time in terms of mixing lengths. In the homogeneous flow
a small fluid element of uniform concentration comes regime, these lengths may range from the bubble
in contact with a phase boundary through which mass diameter to as large as the column height. In part 2 of
transport occurs. Like film theory, where the film this series, mixing lengths, referred to as dispersivities,
thickness is difficult to determine, the exposure time is were determined for the porous frit and the annular
also difficult to determine; however, in some cases, it shear sparger. A simple expression for the dispersion
can be calculated. As suggested in eq 3, the exposure coefficient suggested by Kantak et al. (1994) was used,
time may be taken as the ratio of some characteristic D ) λug. In the case of the porous frit, the dispersivity,
length, λ, which the fluid element must traverse to some based on the superficial gas velocity, was on the order
characteristic velocity, ω. For example, in the case of of the column height, L. However, for the shear sparger,
noninteracting bubbles rising up through a liquid (string the dispersivity was based on the interstitial gas velocity
bubbling flow regime), the exposure time is the mean and was found to be on the order of the column
908 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

Figure 4. Oxygen-based mass-transfer coefficient determined


from the deoxygenation of water versus ug.

diameter, Dc. The reason why the dispersion coefficient Figure 5. Argument of the Higbie correction function versus gas
for the shear sparger scaled with the interstitial velocity velocity.
instead of the superficial velocity was not fully under-
stood; however, it was observed that the circulatory bulk phase Henry law constant, meeting the criteria for
motion set up by the water jet tended to buffer the strongly hydrophobic compounds requires extreme con-
processes of coalescence and breakage, which ultimately ditions. Figure 5 shows the argument of the correction
led to a weak bubble size dependence on the gas velocity. function, f(z), versus the superficial gas velocity for
This weak dependence may suggest that the energy pyrene and PCP. As one can see in the figure for the
dissipated for mixing, coalescence, and breakage is due cases where the conventional Henry law constant is
more to the energy input by the water jet than that of used to determine Shλ, eliminating air-side resistance
the gas. Kantak et al. (1994) reported that the inter- requires extremely low gas velocities. In this regime
stitial gas velocity in the definition of the axial disper- string bubbling flow prevails, and the example cited
sion coefficient makes it independent of the operating earlier applies. However, operating at such low veloci-
parameters and may explain why the dispersion coef- ties is impractical, as the overall removal efficiency
ficient scaled much better with the interstitial gas under these conditions is very small. The situation
velocity. where the conventional Henry law is used applies to
The characteristic velocity is chosen as the superficial processes where the principal driving force for mass
gas velocity, ug. The choice is largely based on the transfer is bulk phase partitioning. However, in the
commonly observed fact that overall volumetric mass- solvent sublation process, the material adsorbed to the
transfer coefficients, KLav, scale well with ug (Shah et bubble-water interface is removed and recovered in the
al., 1982). Typical power dependencies on gas velocity overlying solvent layer. Thus, unlike bubble fraction-
range from 0.6 to 1.7. Data collected from deoxygen- ation where the adsorbed material is not removed from
tation experiments conducted by our group have shown the water (i.e., no sink), the driving force for mass
dependencies similar to those illustrated in Figure 4. transfer in sublation is enhanced. The effective Henry
When eq 3 is combined with correlations for gas holdup law constant, H, which combines bulk phase partition-
and bubble size obtained from the porous frit and the ing and surface adsorption (Hc + 3KA/a) is then used in
shear sparger (part 2 of series), it can be shown that the determination of Shλ. As shown in Figure 5, the
the volumetric mass-transfer coefficients have 1.1 and criteria for neglecting the correction function in eq 5 are
1.3 overall power dependencies in ug, respectively. met for solvent sublation; however, the appropriate
Thus, the choice of ug as the characteristic velocity mixing length must still be used for calculating k. It is
appears to be appropriate. emphasized that Figure 5 applies only to the homoge-
Having defined the exposure time as the character- neous flow regime as the nature of mixing changes
istic mixing length (dispersivity) divided by ug for mass greatly as the flow becomes heterogeneous.
bubbling in bubble columns, the criteria for neglecting Inserting the dispersivities obtained from the porous
air-side resistance in eq 5 may be written in terms of frit (L) and the shear sparger (Dc) into eq 3, one may
the Sherwood and Peclet numbers. Equation 6 defines estimate the relative effectiveness of the shear sparger
over the porous frit

x
DAB

x x
Hcκg g 4DAB πtefrit tefrit
x
te kshear L
) x15 )
kfrit πteshear 4DAB teshear Dc
) ) )
DAB
(Hcκg) g 2 3.87 (7)
λ/ug
In terms of the volumetric mass-transfer coefficients,
(λHcκg) 2 2
λ DAB this ratio becomes
g
DAB2 DAB2(λ/ug) 3
(kav)shear
) 3.87
( a) shear
) 3.87 × 0.503ug0.15 (8)
Shλ g Peλ1/2 (6) (kav)frit
(3a )
frit
the criteria for which air-side resistance may be ne-
glected in terms of the relative degree of backmixing or where the gas velocity power laws determined in part
circulation, Peλ. Given that Shλ is dependent upon the 2 have been substituted for the hydrodynamic proper-
Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997 909
ties. The average ratio is then determined by integrat-
ing eq 8 over the gas velocity limits defining the
homogeneous flow regime,

( ) ∫0.14.0cm/s
cm/s
(kav)shear 1.95ug0.15 dug
) 2.1 (9)
∫0.14.0cm/s
) cm/s
(kav)frit avg dug

Interestingly, this result compares very well with


experimental data (Burns, 1995) collected from the 5.5
in. diameter column described in part 2. The data
indicated that mass-transfer coefficients obtained from
a shear sparger (porous surface area ) 0.05 ft2) are
approximately 2 times greater than those obtained from
a flexible rubber membrane (Flexisparger).
Effect of the Henry Enhancement Factor (Sur- Figure 6. Effect of the Henry enhancement factor on fractional
face Activity). The effectiveness of the sublation removal in solvent sublation and bubble fractionation.
mechanism is largely based upon the surface activity
of the organic compound. This activity can be quantified sublation experiments, R-117 and R-30. For the bubble
in terms of the Henry enhancement factor, H. For fractionation runs, the removal efficiencies for both
example, based on the thermodynamic properties, Hc neutral PCP and pyrene were about 68% or less. For
and KA, cited in part 1 for pyrene and neutral PCP, the the sublation experiments, however, the removal ef-
enhancement factor can be calculated as a function of ficiencies of PCP and pyrene were significantly greater,
the bubble radius, a. In this study, the enhancement 90% and 96%, respectively. Clearly, these data are
factors for both compounds are significantly greater evidence that solvent sublation has a distinct advantage
than 1. However, the enhancement factor for pyrene over bubble fractionation (or diffused aeration). More
is consistently predicted to be 46% greater than that importantly, the data in Figure 6 also confirm the notion
for PCP for the range of bubble radii, 0.25-2.5 mm. This that most conventional processes do not exploit the
implies not only a greater driving force for the mass interface phase. As shown in the figure, the enhance-
transfer of pyrene but also a greater removal efficiency. ment factors for PCP and pyrene under the conditions
As illustrated in Figure 4, the fractional removal of specified were 77 and 117, respectively; and for the
pyrene was greater than that for PCP for all cases. Even sublation experiments, these were good indicators of the
the pyrene data collected from the porous frit showed relative removal efficiencies. However, in the case of
better performance than the PCP data collected from
the bubble fractionation experiments, these enhance-
the shear sparger.
ment factors gave no indication of the relative perfor-
To further illustrate the influence of surface activity,
mance. The reason for this is that in these experiments
run R-16 was conducted in which ionic PCP was
the sorbate at the interface phase remained in the
sublated. As shown in Table 1, the pH of the water was
column.
maintained at 9. The fractional removal for the run was
0.127, which was significantly less than the 0.853 value Model Predictions. Predicting sublation perfor-
obtained from run R-24 which was conducted under mance is a challenging problem given the large number
similar operating conditions but with a pH of 3. The of parameters and the relative dependencies they have
reason for this dependence is that the pH affects the on one another (Smith et al., 1996a). Many parameters,
available amount of the neutral PCP which is suscep- however, are easily obtained such as the column dimen-
tible to the sublation mechanism. Because the pKA of sions, sparger type, fluid properties (e.g., water viscosity,
PCP is 4.74, the degree of dissociation of PCP at pH ) density, surface tension), and operating conditions,
9 is nearly 100%. Consequently, a very low removal while others are obtained through simple calculations
efficiency was observed for the run; however, under and experiments. The values of the parameters used
acidic conditions (R-24), the removal efficiency was in this study are listed in Table 3. The thermodynamic
significantly larger because essentially all of the PCP properties, Hc and KA, appearing in the table were
was surface active. obtained from Montgomery and Welkom (1990) and Hoff
Surface adsorption at an interface is certainly not et al. (1993), respectively. The Kow for neutral PCP was
limited to just solvent sublation. In any interface-mass determined experimentally by extracting three samples
transfer process, surface adsorption may play a role in of an aqueous solution (8 mg/L of PCP) with different
the overall transport. However, whether or not the amounts of mineral oil and measuring the extraction
effect of the enhancement factor is exploited is a efficiencies. The oil-to-water volume ratios used were
completely different issue. For example, in most con- 0.018, 0.030, and 0.050. Plotting the data as FR/(1 -
ventional separation devices such as absorption, strip- FR) versus the oil-to-water volume ratio yielded a slope
ping, and extraction, the phases which are used to equal to Kow (Kow of PCP ) 732). The Kow for pyrene
convect the species of interest are only the bulk phases. was estimated as the ratio of its oil solubility to its water
That is, the interface phase, which is present in each of solubility. Equation 10 (Prausnitz et al., 1986) was used
the devices, is not convected in or out. As such, the to estimate the solubility of pyrene in mineral oil. Table
material which adsorbs to the interface remains in the 4 summarizes the values used in the calculation of the
device. To support this argument, bubble fractionation solubility as well as the value of Kow which results.
runs, R-131 and R-58, were conducted which investi-

[ ( )]
gated the removal in the absence of mineral oil. In Tmelt
1 ∆Hfus
Figure 6, fractional removal results of these experiments x∼ exp 1- (10)
are compared to those obtained from corresponding γ RTmelt T
910 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

Table 3. Model Parameters Used in the Series CSTR The procedure for determining di was to adjust it until
Model (SCM) the fractional removal as predicted by the SCM matched
pyrene pyrene shear neutral PCP the experimental fractional removal to at least three
porous frit sparger shear sparger decimal places. In Table 5, the column labeled “exact
Thermodynamic di” contains those values which gave the best agreement
Hc 7.65 × 10-4 7.65 × 10-4 1.39 × 10-4 between theory and experiment. Interestingly, all of the
KA (cm) 5.39 × 10-3 5.39 × 10-3 6.70 × 10-4 adjusted values are nearly of the same order of magni-
Kow 76412 76412 732 tude, ≈O(10-3) cm, regardless of compound type and
Design sparger type. This observation is a good indication that
Dc (cm) 10.16 10.16 10.16 the transport mechanisms in sublation have been ac-
L (cm) 152.4 152.4 152.4 counted for appropriately. On closer inspection, how-
Hydrodynamic ever, the corresponding wake-to-bubble volume ratios,
a (cm) 0.122ug0.573 0.174ug0.24 0.174ug0.24 Vw/Vb ≈ 3di/a, are only on the order of ≈O(10-2). Fan
 0.0491ug1.21 a 0.0352ug1.031 0.0352ug1.031 and Tsuchyia (1990) have compiled wake-to-bubble
λ (cm) 152.4 10.16 10.16
volume data from numerous investigators which show
Kinetic that, for spherically-capped bubbles with effective di-
no. of CSTR stages n ) 1 (i ) 2) n ) 1 (i ) 2) n ) 1 (i ) 2)
ameters of 4-20 mm, Vw/Vb is on the order of 5. The
DAB (cm2/s) 6.70 × 10-6 6.70 × 10-6 7.19 × 10-6
bubbles observed in the sublation experiments, however,
x x x
k, Higbie model 4DAB 4DAB 4DAB
(cm/s) were much more spherically-shaped and smaller. Thus,
πL/ug πDc/ug πDc/ug the wakes associated with these bubbles should be very
kl, film theory 10-4 10-4 10-4 small and most likely are the results of entirely different
(cm/s) physics than those discussed by Fan and Tsuchyia.
a The power-law model here varies slightly from that cited in Based on the relative magnitude of di/a, it is believed
part 2 as it is based only on the data obtained from the porous that di is related to the viscous boundary layer that
frit and not the flexible rubber sparger. develops around the bubble.
In the development of the velocity distribution for a
Table 4. Values Used To Calculate the Solubility of rising gas bubble in a liquid, Levich (1962) found that
Pyrene in Mineral Oil and the Oil-Water Partition
Coefficient the boundary layer thickness may be estimated from
the bubble radius, the kinematic viscosity, and the rise
T (K) 298 velocity, U, as shown in eq 11. (The same result is
Tmelt (K) 424
∆Hfus (estimated) (J/g) 127.3
µa 1/2
γ pyrene in oil (UNIFAC)
water solubility (mol frac)
3.4
1.20 × 10-8
di ∼ (FU ) (11)
oil solubility (mol frac) 0.0134
Kow (lt/lt) 76412 predicted by Boundary Layer Theory (Bird et al., 1960).)
Thus, substituting typical values into the equation, one
The liquid diffusivities, DAB, were estimated from the can estimate the boundary layer around a rising bubble,

( )
Wilke-Chang correlation (Geankoplis, 1993), and the
hydrodynamic correlations shown in the table are based 0.01 g/cm‚s × 0.1 cm 1/2
di ∼ ) 5.8 × 10-3 cm (12)
on the data obtained in part 2. Of all the SCM 1 g/cm3 × 30 cm/s
parameters, the amount of wake (water), di, which is
entrained into the solvent is the most difficult to Interestingly, this prediction of di is consistent with
determine and is therefore chosen as the single regres- those shown in Table 5, indicating that the water which
sion variable. is entrained into the solvent is indeed the result of the

Table 5. Adjusted and Regressed Bubble-Wake Thicknesses Used To Evaluate the Theoretical Fractional Removal
SCM (i ) 2)
bubble
Reynolds no. di (cm) FR relative
sublation run (FUa)/µ sparger a (cm) exact regressed regressed experimental error (%)
R-73-9-21-94 121 porous frit 0.048 2.45 × 10-3 1.95 × 10-3 0.779 0.809 -3.75
R-79-9-27-94 121 porous frit 0.048 2.43 × 10-3 1.95 × 10-3 0.841 0.863 -2.60
R-80-9-29-94 121 porous frit 0.048 2.21 × 10-3 1.95 × 10-3 0.779 0.796 -2.17
R-87-10-12-94 121 porous frit 0.048 2.40 × 10-3 1.95 × 10-3 0.840 0.862 -2.49
R-88-10-13-94 517 porous frit 0.078 1.83 × 10-3 1.38 × 10-3 0.873 0.892 -2.14
R-89-10-15-94 345 porous frit 0.115 8.26 × 10-4 1.03 × 10-3 0.904 0.896 0.93
R-92-11-1-94 32 porous frit 0.031 1.87 × 10-3 2.61 × 10-3 0.818 0.771 6.16
R-94-11-15-94 32 porous frit 0.031 2.49 × 10-3 2.61 × 10-3 0.693 0.684 1.28
R-97-11-17-94 192 porous frit 0.064 8.59 × 10-4 1.58 × 10-3 0.924 0.893 3.51
R-99-11-30-94 234 porous frit 0.078 1.58 × 10-3 1.38 × 10-3 0.821 0.833 -1.45
R-100-12-1-94 498 porous frit 0.166 7.76 × 10-4 7.85 × 10-4 0.901 0.901 0.03
R-106-1-18-95 432 annular shear 0.144 2.69 × 10-3 2.83 × 10-3 0.890 0.887 0.32
R-113-2-9-95 501 annular shear 0.167 4.07 × 10-3 2.29 × 10-3 0.926 0.945 -2.01
R-117-3-7-95 549 annular shear 0.183 4.49 × 10-3 2.00 × 10-3 0.943 0.962 -1.94
R-120-3-21-95 534 annular shear 0.178 6.93 × 10-4 2.08 × 10-3 0.968 0.956 1.30
R-124-4-6-95 432 annular shear 0.144 6.99 × 10-3 2.83 × 10-3 0.802 0.882 -9.12
R-24-9-20-95 519 annular shear 0.173 1.55 × 10-3 2.20 × 10-3 0.880 0.853 3.13
R-30-9-26-95 573 annular shear 0.191 2.20 × 10-3 1.90 × 10-3 0.898 0.907 -0.90
R-33-9-29-95 573 annular shear 0.191 1.66 × 10-3 1.90 × 10-3 0.947 0.942 0.46
R-37-10-3-95 459 annular shear 0.153 2.60 × 10-3 2.63 × 10-3 0.745 0.744 0.12
R-40-10-9-95 459 annular shear 0.153 2.60 × 10-3 2.63 × 10-3 0.796 0.783 1.57
R-43-10-10-95 480 annular shear 0.160 2.60 × 10-3 2.46 × 10-3 0.864 0.858 0.77
Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997 911
eliminate the bubble radius, the wake ratio then be-
comes a function of the superficial gas velocity

( ) ()
Vw
Vb frit

di
a frit
∝ ug-3/4

( ) ()
Vw
Vb shear

di
a shear
∝ ug-3/8 (15)

Like most hydrodynamic properties, eq 15 suggests that


the appropriate variable to regress against is ug. Fur-
thermore, it predicts that the wake ratio for the shear
sparger has a weaker dependence on ug than that for
the porous frit. As shown in eq 16, this was indeed the
case; however, the actual dependencies are somewhat
Figure 7. Frequency distribution of measured water droplets larger than the predicted values. Moreover, the R2 for
entrained in the mineral oil during sublation. the shear sparger was only 43%, indicating perhaps that
di/a is independent of gas velocity or that there is an
boundary layer and not the wakesat least not the wake insufficient amount of data on which to base a conclu-
defined in the conventional sense. To seek further sion.
confirmation, attempts to measure the sizes of the water

()
droplets in the mineral oil were made with the JAVA di
video image analyzer described in part 2 and a 10 power ) 0.024ug-0.955 R2 ) 94%
Olympus video camera. Although measuring the drop- a frit

()
let sizes was not difficult, capturing still images of the
di
droplets was. The reasons for this were 2-fold. The first ) 0.038ug-0.600 R2 ) 43% (16)
was that the fast speed at which the droplets traversed a shear
the field of view caused streaking in many of the images.
The second problem was that the mineral oil became When the entire data set is used in the regression, the
opaque due to aeration and had to be frequently dependence on ug is ≈-3/4, and the correlation coefficient
replaced. Nonetheless, some still images were captured is 84%. This at least tends to confirm the functional
and analyzed. A histogram of the droplet sizes captured form of eq 15.
at QA ) 1592 cm3/min and Qw) 0 is shown in Figure 7. The regressed di’s which result from eq 16 are shown
The weighted average for the data is 0.13 cm. When in Table 5. These values were used in the SCM to
the values of di and a in Table 5 (for example, run R-106) calculate the removal efficiencies. As expected, the
are used to calculate the equivalent water droplet regressed values do introduce error between the pre-
diameter, d̃, that would result in the oil phase, one dicted and experimental removal efficiencies; however,
discovers that d̃ is, in fact, nearly the same as the the error in most cases is less than (3%. The largest
measured value. deviation between theory and experiment was 9.12%
and occurred for run R-124. In this particular run,
d̃ ) 2[(a + di)3 - a3]1/3 ) however, the steady-state value was suspect due to
interface control problems and may explain the large
2[(0.144 + 2.69 × 10-3)3 - 0.1443]1/3 ) 0.11 cm (13) error. A better way to assess the model predictions is
to consider the parity plot in Figure 8a. The absence
Thus, it is reasonable to conclude that bubble wake of biases among the data sets is a strong indication of
entrainment into the solvent is dictated by the viscous model validity. Moreover, this observation suggests
boundary layer. It is important, though, to emphasize that a generalized correlation of the type described in
that this finding is only applicable to the homogeneous part 1 may be sought. For cases where the water-
flow regime. At higher gas velocities the dynamics solvent capacity factor, Pw ≡ Qw/Kow/Qo, is sufficiently
governing wake phenomena are vastly different. Though small (Pw e 0.013), it was found that the fractional
some light has been shed on the physics of the wake removal may be expressed in terms of the water-side
entrainment mechanism, the issue of determining the entrainment number, Ew, and the air-water Stanton
proper variable with which to regress di still remains. number, StwA. (Ew describes the mass transfer associ-
Levich (1962) showed that, for bubble Reynolds ated with the entrainment of the bubble wake into the
numbers, (FUa/µ), < 600, the rise velocity depended solvent, and StwA describes the mass transfer from the
upon the square of the bubble radius. Experimental water phase to the air bubbles.) Since the PCP-
data (Clift et al., 1978) confirm this type of dependence; mineral oil system does not meet the condition, the
however, for contaminated water the dependence has following correlation is valid only for the pyrene data
been shown to be much more linear. Substituting a2
for U and rearranging eq 11, one finds that the theoreti- FR
cal dependence of the wake-to-bubble volume ratio is ) 3.42E0.24 0.39
w StwA R2 ) 93% (17)
1 - FR
Vw di where FR is the fractional removal. Equation 17
∝ ∝ a-3/2 (14)
Vb a compares well with the theoretically-based correlation
derived in part 1 in that the mass transfer associated
(An a-1 dependence results if the contaminated water with the air bubbles (StwA) is weighted more than that
dependence is used; however, the subtle differences associated with entrainment (Ew). The values of the
between the two do not affect the conclusion.) When exponents appearing in eq 17 and the leading coefficient,
the correlations in Table 3 for bubble size are used to however, do differ from those in the theoretical correla-
912 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

analysis, may be compared to that recovered in the oil.


As shown in the third column of Table 6, the mass
recovered was less than what was removed in all but
two cases (actually four; two are within experimental
error). This suggests that, in general, there were
unaccounted losses. Though one possibility is that some
of the pyrene may have been lost to the atmosphere, it
is highly unlikely that this would occur given the low
volatility of pyrene. The most probable loss mechanism
is due to the fact that, during the course of the sublation
experiments, small droplets of oil were continuously
sprayed onto and, to a limited extent, over the walls of
the overflow reservoir. The oil which was sprayed on
the walls formed a film of residue which remained until
it was wiped clean at the end of a run. These oil losses
were estimated to be on the order of 3% of the total oil
processed and can account for as much as a 10% mass
deficiency. For those runs in which the mass balances
disagree by more than 10%, the discrepancies may in
part be attributed to the fact that the variation between
the feed stock concentrations was as much as 15 µg/L.
Since the volumes of the feed stocks fed to the process
were not necessarily equal, calculations involving the
average feed water concentration may result in errors
greater than 10%.
An important measure of the utility of the solvent
phase in any liquid-liquid type separation process is
the separation factor. It is defined as the ratio of the
solvent effluent concentration to the water effluent
concentration. As shown in Table 6, the separation
factors determined from the pyrene sublation data
Figure 8. (a) Parity plot of the theoretical fractional removal
calculated by the SCM versus experimental data. (b) Experimen-
ranged from about 30 to well over 300. Such large
tally determined separation factors (corrected for dilution due to values are unique to sublation and are indicative of its
startup inventory) versus the theoretical separation factor. superiority over solvent extraction. For example, in
solvent extraction, the ratio of the solvent flow rate to
tion cited in part 1. Nonetheless, the correlation does the water flow rate is generally much greater than the
confirm that, for strongly hydrophobic compounds, ratios used in this study. Consequently, the separation
sublation performance may be estimated quickly and factor for the corresponding liquid-liquid extraction
accurately by considering just the two important mass- process would be much less than those shown in Table
transfer mechanisms associated with rising air bubbles. 6. Also appearing in the table are the theoretical oil
Overall Mass Balance and the Separation Fac- concentrations predicted by the SCM. A comparison of
tor. One of the concerns with past sublation studies these values with the experimentally determined con-
has been the lack of an overall mass balance, which centrations indicates that, in all but one case, the theory
shows that the sublate is recovered in the solvent phase. overpredicts the concentration. The reason for this is
Thus, for the pyrene experiments conducted in this attributed to the volume of oil used to fill the overflow
study, the total amount of oil collected from the experi- reservoir during startup. The initial oil loadings were
ments was analyzed at the end of each run. In this way a significant fraction (25-46%) of the total oil collected.
the mass of pyrene removed, based on the water Correcting the oil concentrations for dilution by the

Table 6. Overall Mass Balance and Separation Factor Data for the Sublation of Pyrene
effluent oil separation
mass removed average feed average effluent conc. (Co) effluent oil factor
pyrene (water analysis) mass recovered water conc. water conc. experiment conc. (Co) theory (CoFoil/Cw)
sublation run (g) (oil analysis) (g) % error (Cwo) (µg/L) (Cw) (µg/L) (ppm (wt)) (ppm (wt)) experiment
R-73-9-21-94 5.59 × 10-3 5.52 × 10-3 -1.2 96 18.3 0.817 0.908 38
R-79-9-27-94 8.05 × 10-3 6.42 × 10-3 -20.2 119 16.3 0.543 0.811 29
R-80-9-29-94 7.34 × 10-3 5.99 × 10-3 -18.4 119 24.3 0.742 1.126 26
R-87-10-12-94 3.62 × 10-3 3.63 × 10-3 0.3 109 15.0 1.120 1.486 64
R-88-10-13-94 6.98 × 10-3 6.15 × 10-3 -11.9 126 13.6 1.030 1.784 65
R-89-10-15-94 4.78 × 10-3 4.55 × 10-3 -4.8 127 13.2 0.960 1.863 63
R-92-11-1-94 3.58 × 10-3 3.52 × 10-3 -1.6 100 22.9 0.944 1.328 35
R-94-11-15-94 5.52 × 10-3 5.25 × 10-3 -4.9 106 33.5 1.170 2.382 30
R-97-11-17-94 1.70 × 10-3 2.04 × 10-3 20.2 97 10.4 0.622 0.725 52
R-99-11-30-94 8.18 × 10-3 6.75 × 10-3 -17.5 120 20.0 1.220 2.394 52
R-100-12-1-94 5.06 × 10-3 5.04 × 10-3 -0.4 106 10.5 1.570 2.322 129
R-106-1-18-95 4.59 × 10-3 4.22 × 10-3 -8.0 92 10.4 0.970 1.333 80
R-113-2-9-95 4.67 × 10-3 4.27 × 10-3 -8.6 103 5.7 0.840 1.555 128
R-117-3-7-95 4.81 × 10-3 5.60 × 10-3 16.3 96 3.6 1.370 1.476 323
R-120-3-21-95 2.32 × 10-3 3.19 × 10-3 37.5 106 4.7 0.870 0.832 160
R-124-4-6-95 1.01 × 10-2 1.00 × 10-2 -0.2 120 14.2 1.860 3.142 113
Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997 913
initial loadings greatly improves the agreement, as SEARCHER, nor any person acting on behalf of any of
shown by Figure 8b. them (a) makes any warranty or representation, ex-
pressed or implied, with respect to the accuracy, com-
Conclusion pleteness, or usefulness of the information contained in
this report, or that the use of any information, ap-
Three-phase, continuous solvent sublation of pyrene paratus, method, or process disclosed in this report may
and pentachlorophenol (PCP) was successfully demon- not infringe on privately-owned rights or (b) assumes
strated. The conductivity-based controller proved to be any liability with respect to the use of, or for damages
an effective and reliable method for controlling the resulting from the use of, any information, apparatus,
height of the solvent-water interface and was superior method, or process disclosed in this report. We give
to ultrasonic level control. The benefits of the annular special thanks to Mike Kurtz for his valuable assistance
shear sparger were realized as steady-state removal with the model-based control algorithm.
efficiencies as high as 96% for pyrene and 94% for PCP
were observed. Nomenclature (also applies to parts 1 and 2 of
Sublation results also supported that for hydrophobic the series)
organic compounds the effect of the solvent flow rate is
very weak, and as model predictions suggested, only an a ) bubble radius, cm
amount of solvent which ensures a steady state is A ) dimensionless air concentration, A ≡ CA/HCow
necessary. This unique feature of solvent sublation had CA ) effective concentration of pollutant in air, mol/cm3
a profound effect on the separation factor, which is a Co ) concentration of pollutant in solvent, mol/cm3
measure of solvent utility. Separation factors as high Cw ) concentration of pollutant in water, mol/cm3
as 300 were calculated, which demonstrated the supe- CP ) chlorinated phenol
riority of solvent sublation over solvent extraction. di ) bubble-wake thickness, cm
d̃ ) water droplet diameter, cm
It was shown that the Henry enhancement factor,
D ) dispersion coefficient, cm2/s
which is a measure of the additional air-water parti-
Dc ) column diameter, cm
tioning that arises from surface adsorption, is an DAB ) binary diffusivity, cm2/s
excellent indicator of a compound’s susceptibility to Ew ) water-side entrainment number, Ew ≡ 3QAdi/Qwa
solvent sublation. For example, removal efficiencies
FR ) fractional removal, FR ≡ 1 - Cw/Cow ) 1 - Aeff/Ainf
obtained for pyrene, which has an enhancement factor
H ) effective Henry law constant, Hc + 3KA/a
of approximately 117, were consistently greater than
Hc ) Henry law constant, cm3/cm3
those obtained for PCP, which has an enhancement ∆Hfus ) heat of fusion, J/mol
factor of only 77. It was further shown that for i ) number of stages comprising the SCM
processes which do not exploit the surface activity of k ) water-side mass-transfer coefficient (air/water), cm/s
the air-water interface, such as bubble fractionation, kl ) overall water-side mass-transfer coefficient (solvent/
the removal efficiencies for pyrene and PCP were water), cm/s
limited to 68%. KLav ) overall volumetric mass-transfer coefficient, s-1
An analysis of the overall resistance to mass transfer KA ) surface adsorption constant, cm
showed that, for nonvolatile hydrophobic compounds, Kow ) solvent-water partition coefficient, cm3/cm3
gas-side resistance is dominant in conventional separa- L ) length of bubble column, cm
tion processes. However, in solvent sublation where the n ) number of stages comprising the water column, n ) i
surface-active nature of hydrophobic compounds is -1
exploited, gas-side resistance is circumvented. Accord- Pw ) water-solvent capacity factor, Pw ≡ Qw/QoKow
ingly, a modified Higbie penetration model of the overall PAH ) polynuclear aromatic hydrocarbon
mass-transfer coefficient was developed as a function QA ) air flow rate, cm3/s (cm3/min)
of the characteristic mixing length, λ, and characteristic Qo ) solvent flow rate, cm3/s (cm3/min)
velocity, ω. It was argued that λ is the dispersivity Qw ) water flow rate, cm3/s (cm3/min)
determined by axial dispersion experiments, and ω (for R ) gas constant, 8.314 J/mol‚K
conditions of mass bubbling in the homogeneous flow S ) dimensionless solvent concentration, S ≡ Co/KowCow
regime) is the superficial gas velocity. Under this Stw ) water-side Stanton number (solvent/water), Stw ≡
criteria, Higbie model predictions showed that the use klπrc2(1 - )/Qw
of the annular shear sparger improved mass transfer StwA ) water-side Stanton number (air/water), StwA ≡
by a factor of 2 over the fine porous frit. Agreement 3kπrc2L/aQw
te ) exposure time
between experimental data and model predictions sup-
T ) temperature, K
ported this assertion.
Tmelt ) melting point, K
Lastly, the Series CSTR Model (SCM) proved to be ug ) superficial gas velocity, cm/s
an excellent tool for predicting fractional removal, as U ) bubble rise velocity, cm/s
most predictions were within (3%. For the sublation Vb ) bubble volume, cm3
of pyrene, fractional removal data were correlated in Vo ) volume of solvent, cm3
terms of the air-water Stanton number, StwA, and the Vw ) wake volume, cm3
water-side entrainment number, Ew. The correlation W ) dimensionless water concentration, W ≡ Cw/Cow
supported the idea that the two most important mass- x ) solubility, mole fraction
transfer mechanisms are air-water mass transfer and
bubble-wake entrainment. Greek Letters
 ) gas holdup, cm3/cm3
Acknowledgment H ) Henry enhancement factor, H ≡ 1 + 3KA/aHc
γ ) activity coefficient
This paper was prepared by Louisiana State Univer- κl ) liquid-side mass-transfer coefficient, cm/s
sity as an account of work sponsored by AIChE/CWRT. κg ) gas-side mass-transfer coefficient, cm/s
Neither AIChE/CWRT, members of AIChE/CWRT, RE- λ ) characteristic length or dispersivity, cm
914 Ind. Eng. Chem. Res., Vol. 36, No. 3, 1997

µ ) fluid viscosity, g/cm‚s Levich, V. G. Physicochemical Hydrodynamics; Prentice-Hall:


F ) fluid density, g/cm3 Englewood Cliffs, NJ, 1962; Chapter 8.
σ ) surface tension, dyn/cm Montgomery, J. H.; Welkom, L. M. Groundwater Chemicals Desk
Σ ) separation factor, Σ ≡ CoKowCw Reference, 2 ed.; Lewis Publishers, Inc.: Chelsea, U.K., 1990; p
640.
ω ) characteristic velocity, cm/s Prausnitz, J. M.; Lichtenthaler, R. N.; Azevedo, E. G. Molecular
Thermodynamics of Fluid-Phase Equilibria; Prentice-Hall: En-
Literature Cited glewood Cliffs, NJ, 1986; Chapter 9.
Shah, Y. T.; Kelkar, B. G.; Godbole, S. P.; Deckwer, W. D. Design
Baird, M. H. I.; Rice, R. G. Axial Dispersion in Large Unbaffled Parameters Estimations for Bubble Column Reactors. AIChE
Columns. J. Chem. Eng. 1975, 9, 171-174. J. 1982, 28, 353-379.
Bird, R. B.; Stewart, W. E.; Lightfoot, E. N. Transport Phenomena; Sherwood, T. K.; Pigford, R. L.; Wilke, C. R. Mass Transfer;
John Wiley & Sons: New York, 1960; Chapter 4. McGraw-Hill: New York, 1975; Chapter 5.
Burns, L. F. The Effect of Reduced Surface Tension on Mass Smith, J. S.; Valsaraj, K. T.; Thibodeaux, L. J. Bubble Column
Transfer and Fluid Dynamics in Bubble Column Reactors. M.S. Reactors for Wastewater Treatment. 1. Theory and Modeling
Thesis, Louisiana State University, Baton Rouge, LA, 1995. of Continuous-Countercurrent Solvent Sublation. Ind. Eng.
Chern, J. M.; Yu, C. F. Volatile Organic Compound Emission Rate Chem. Res. 1996a, 35, 1688-1699.
from Diffused Aeration Systems. 1. Mass Transfer Modeling. Smith, J. S.; Burns, L. F.; Valsaraj, K. T.; Thibodeaux, L. J. Bubble
Ind. Eng. Chem. Res. 1995, 34, 2634-2643. Column Reactors for Wastewater Treatment. 2. The Effect of
Clift, R.; Grace, J. R.; Weber, M. E. Bubbles, Drops, and Particles; Sparger Design on Sublation Column Hydrodynamics in the
Academic Press: San Diego, 1978; Chapter 7. Homogeneous Flow Regime. Ind. Eng. Chem. Res. 1996b, 35,
Cusack, R. W. Solve Wastewater Problems with Liquid/Liquid 1700-1710.
Extraction. Chem. Eng. Prog. 1996, 4, 56-63. Taylor, R.; Krishna, R. Multicomponent Mass Transfer; Wiley:
Fan, L. S.; Tsuchiya, K. Bubble Wake Dynamics in Liquids and New York, 1993; Chapter 9.
Liquid-Solid Suspensions; Brenner, H., Ed.; Butterworth-
Heinemann: Stoneham, MA, 1990; Chapter 5. Received for review August 22, 1996
Geankoplis, C. J. Transport Processes and Unit Operations; Revised manuscript received December 12, 1996
Prentice-Hall: Englewood Cliffs, NJ, 1993; Chapter 6. Accepted December 18, 1996X
Hoff, J. T.; Mackay, D.; Gillham, R.; Shiu, W. Y. Partitioning of
Organic Chemicals at the Air-Water Interface in Environmen- IE9605241
tal Systems. Environ. Sci. Technol. 1993, 27, 2174-2180.
Kantak, M. V.; Shetty, S. A.; Kelkar, B. G. Liquid Phase Back-
X Abstract published in Advance ACS Abstracts, February
mixing in Bubble Column ReactorssA New Correlation. Chem.
Eng. Commun. 1994, 127, 23-34. 1, 1997.

You might also like