You are on page 1of 12

+ +

1688 Ind. Eng. Chem. Res. 1996, 35, 1688-1699

GENERAL RESEARCH

Bubble Column Reactors for Wastewater Treatment. 1. Theory and


Modeling of Continuous Countercurrent Solvent Sublation
Jeffrey S. Smith, Kalliat T. Valsaraj,* and Louis J. Thibodeaux
Department of Chemical Engineering, Louisiana State University, Baton Rouge, Louisiana 70803

Solvent sublation is a nonfoaming wastewater treatment process that combines the benefits of
bubble fractionation and liquid-liquid extraction in a way that does not require mixers, settlers,
or subsequent downstream treatment. A review of past work on small lab-scale batch columns
revealed that removal efficiencies of nonvolatile and volatile organic compounds are generally
higher than those observed in bubble fractionation, air stripping, and conventional liquid-liquid
extraction. In this work, the first of a three-part series, the transport mechanisms in a three-
phase continuous, countercurrent sublation process are presented. Two mathematical models,
namely, the series CSTR model (SCM) and the two-phase axial dispersion model (ADM2), are
developed. It is shown that these two models are equivalent and can be used interchangeably
with the aid of a simple expression. Nondimensional correlations, based upon simulated data
obtained from the SCM, are generated to predict the steady-state fractional removal, FR, and
the separation factor, Σ (ratio of effluent solvent concentration to effluent water concentration),
for strongly hydrophobic compounds. The effects of operational, hydrodynamic, thermodynamic,
and design variables on sublation performance are discussed.

Introduction surround each globule. CGAs (25-150 µm) are similar


except that they contain gas inside the soap film instead
Processes which exploit the tendency for organic
of oil. In the water column, the polyaphrons adhere to
compounds, ions, or particles to adsorb or to attach to
the surfaces of the CGAs. As the CGAs rise, solute
air-water interfaces of rising bubbles are generally
referred to as “adsorptive bubble processes” (Sebba, molecules in the aqueous phase partition themselves
1962; Karger et al., 1967b; Karger, 1972). There are between the polyaphrons and the water. At the surface
two types of adsorptive bubble processes, foaming and of the column, the polyaphrons coalesce forming a layer
nonfoaming. In foaming processes, surfactants are of solvent containing the solute. In principle, this
introduced to generate a froth or foam for the removal technique is an effective means of removing organic
of particles by flotation. In some processes, the particles compounds from water; however, it removes compounds
which are removed are sufficiently hydrophobic that at the expense of contaminating the water with residual
surfactant is not necessary. Important applications solvent and surfactant. Therefore, the effluent water
include the cleaning and recovery of coal and graphite from a predispersed solvent extraction column must
as well as valuable metals from ores and minerals undergo further treatment (Valsaraj, 1995).
(Finch and Dobby, 1990). In nonfoaming processes, A process related to predispersed solvent extraction
surfactant may be used, but not under conditions which is solvent sublation (Sebba, 1962; Karger, 1972). In a
produce a foam or a froth for flotation purposes. Non- review paper (Valsaraj et al., 1991), the authors have
foaming processes are better suited for wastewater described the transport mechanisms that are important
treatment since procedures to collapse and dispose wet to solvent sublation as well as discussed the general
foams are not required (Valsaraj, 1995). The classic
“state-of-the-art”. The basic advantage of solvent sub-
example of a nonfoaming adsorptive process is bubble
lation over predispersed solvent extraction and conven-
fractionation (Lemlich, 1966; Kown and Wang, 1971).
Bubble fractionation is basically air-stripping in a tional solvent extraction is that intimate contact be-
bubble column. Other examples of nonfoaming proc- tween the organic solvent and the aqueous phase is
esses are predispersed solvent extraction and solvent prevented. This feature overcomes the problem of
sublation (Valsaraj, 1995). These processes are unique residual solvent in the aqueous phase. Furthermore,
in that they blend the benefits of bubble fractionation the use of a surfactant is not crucial to the operation.
with other separation processes. There have been a number of examples where the
In predispersed solvent extraction, gas (air) and hydrophobic nature of neutral-nonpolar compounds has
solvent (oil) are mixed with surfactant to generate been utilized to remove them in solvent sublation. Some
polyaphrons and colloidal gas aphrons (CGAs) (Mich- examples appear in Table 1a in which the solute
elsen et al., 1986; Caballero et al., 1989). Polyaphrons removed, solvent, and the operating conditions are
are globules of oil (95%) and water (5%) that range from listed. A surfactant is only necessary when the com-
submicron to approximately 50 µm in size. They are pound to be removed is ionizable, e.g., ionic dyes, metal
stabilized in aqueous solutions by thin soap films which complexes (Cervera et al., 1982; Huiru and Xiuyu, 1988;
Wang et al., 1993), or phenols, as illustrated in Table
* To whom correspondence should be addressed. 1b. In the cases of weakly ionizable species, such as
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1689


Table 1. Representative Solvent Sublation Studies
air flow rate /
surfactant (or column water flow rate
compound complex agent) solvent dimensions max (mL/min) investigators

(a) Neutral Compounds


1,1,1-trichloroethane 1-octanol 5.2 × 116 cm 120/nil Lionel et al., 1981
dichlorobenzenes, mineral oil 3.5 × 85 cm 97/nil Valsaraj and Wilson,
nitrophenols, PCBs, 1983
lindane, endrin
naphthalene, phenanthrene mineral oil 3.5 × 85 cm 97/nil Huang et al., 1983
alkyl phthalates mineral oil 3.6 × 90 cm 180/nil Tamamushi and Wilson,
1984-5
indene, aldrin mineral oil 3.1 × 90 cm 70/nil Foltz et al., 1986
pentachlorophenol (PCP) mineral oil, hexane, 5 × 100 cm 168/nil Valsaraj and Springer,
(pH ) 2.5) 1-octanol 1986
diphenyl mineral oil 3.2 × 100 cm 100/nil Wang and Huang, 1988
o-dichlorobenzene, mineral oil, 8 × 50 cm 133/nil Shin and Coughlin, 1990
toluene 1-octanol,
2-octanol
hexachlorobutadiene, mineral oil 3.5 × 60 cm 180/nil Shih et al., 1990
2,4,6-trichlorophenol
(TCP)
PCP, TCP, naphthalene mineral oil 5 × 100 cm, 250/25 Valsaraj et al., 1992
15 cm ×
150 cm
(b) Ionic Compounds
methyl orange hexadecyltrimethyl- 2-octanol 4.5 × 46 cm 30/nil Karger et al., 1967a
Rhodamine B ammonium bromide,
HTMAB
methyl orange HTMAB 2-octanol 9 cm diam. 167/nil Karger et al., 1970
(2 lt.)
hexacyanoferrate(II) dodecylpyridinium 2-octanol 9 cm diam. 167/nil Spargo and Pinfold, 1970
chloride (2 lt.)
thulium and americium citric acid N/A N/A N/A Stachurski and Szeglowski,
1974
methylene blue and HTMAB 2-octanol 3.5 × 87 cm 57/nil Womack et al., 1982
orange
magneta (cationic dye) sodium lauryl mineral oil 3.5 × 60 cm 120/nil Sheu and Huang, 1987
sulfate
eleven phenolic HTMAB, diisopropyl ether, N/A 60/nil Caballero et al., 1988
compounds stearylamine isopropyl alcohol
phenols ethylhexadecyl- MIBK 6 × 76 cm 65/nil Nolan and McTernan, 1988
ammonium
bromide
neutral and ionic PCP HTMAB mineral oil, decyl 2.3 × 100 cm 30/nil Valsaraj and Thibodeaux,
alcohol 1991a
neutral and Ionic, PCP, HTMAB mineral oil, decyl 2.3 × 100 cm 72/6 1 run with Valsaraj and Thibodeaux,
1,2,4-trichlorobenzene, alcohol 1.8 mL/ min 1991b; Lu et al., 1991
2,3,6-trichloroanisole, solvent flow
TCP rate
Gold HTMAB 2-octanol, MIBK, 3.7 × 40 cm 45/nil Wang et al., 1993
n-butyl acetate,
tributyl phosphate

phenols, the use of surfactant can be avoided by adjust- needed, which makes sublation less expensive. The
ing the pH (Valsaraj and Springer, 1986). biggest advantage over both conventional extraction
Solvent sublation may be thought of as a combination (single-staged) and bubble fractionation is that a higher
of conventional liquid-liquid extraction and bubble degree of fractional removal is possible under certain
fractionation. In Figure 1, countercurrent configura- circumstances (Valsaraj et al., 1991).
tions of these processes are compared. Like all adsorp- As can be seen in Table 1a,b, most of the sublation
tive bubble processes, solvent sublation is similar to work to date has been conducted on small-scale labora-
bubble fractionation. The difference is that, in subla- tory columns operating in a batch mode relative to the
tion, a nonvolatile organic solvent is floated upon the liquid phase. Though this type of information is im-
water column. In bubble fractionation, nonvolatile portant for studying the transport mechanisms in
compounds become enriched in the top portion of the sublation, it is not useful for ascertaining scaleup or
aqueous phase, and volatile organic compounds (VOCs) industrial feasibility.
are lost to the atmosphere unless some other gas-phase Recently (Lu et al., 1991; Valsaraj and Thibodeaux,
treatment is used. In solvent sublation, on the other 1991a,b; Valsaraj et al., 1992), our group investigated
hand, volatile and nonvolatile compounds are recovered semicontinuous operation (continuous, countercurrent
in the solvent phase, just like they are recovered in the air and water phases) and examined the effect of column
solvent phase used in the extraction process. In the diameter (15 versus 5 cm). These studies demonstrated
extraction process, multiple stages are required for any the feasibility of operating continuously and showed
appreciable recovery; however, in sublation, only one that solvent sublation consistently outperformed bubble
stage of solvent-water contact is required. Further- fractionation in terms of fractional removal. Further-
more, in solvent sublation, mixers and settlers are not more, sublation performance was found to be less
+ +

1690 Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996

Figure 1. Comparison of liquid-liquid extraction, bubble fractionation, and solvent sublation.

Table 2. Thermodynamic Properties of Hydrophobic ematical models to predict the steady-state pollutant
Compoundsa concentrations in all three phases of a continuous
PAH (pyrene) CP (PCP) sublation column. We examine the effects of the
(MW ) 202) (MW ) 266) operating variables, thermodynamic variables, and
aqueous solubility (mg/L) 0.135 14 design variables on sublation performance. Model-
vapor pressure (mmHg) 6.80 × 10-7 1.10 × 10-4 generated, fractional removal data are correlated in
log Kow′′ 4.88 5.01 terms of key dimensionless numbers so that the use of
Henry’s constant, Hc (cm3/cm3) 7.65 × 10-4 1.39 × 10-4 the complex analytical solutions may be avoided. In
Interfacial partition constant, 5.39 × 10-3 9.13 × 10-5 part 2, we present hydrodynamic data, such as bubble
KA (cm)b
size and gas holdup, for two types of gas spargers used
a Montgomery and Welkom, 1990. b Mackay et al., 1990; Hoff
in a pilot-scale sublation column. We also show that
et al., 1993. the homogeneous or bubbly flow regime (Shah et al.,
dependent on the column diameter. This was in con- 1982) is the appropriate operating regime for solvent
trast to the bubble fractionation results where the sublation. In part 3 of this series, we will present
performance suffered as the diameter increased. It is steady-state fractional removal data collected from a
generally known that increasing the column diameter continuous, pilot sublation column and a bubble frac-
adversely affects the axial mixing in bubble fraction- tionation column.
ation (Kown and Wang, 1971; Deckwer et al., 1974;
Shah et al., 1982); however, it was concluded that, once Mass Transport Mechanisms
a solvent layer captures the solute as in sublation, it is
retained as the solvent possesses a greater affinity for In solvent sublation, there are three principle path-
the solute. From preliminary studies using the 15 cm ways or transport mechanisms available by which a
diameter column, it was recognized that the inability pollutant may be removed from the aqueous phase. They
to produce small bubbles (<1 mm) at high air flow rates are transport by air bubbles, water entrainment due to
(>600 mL/min) can be a problem in scaled-up operation. the wakes of rising air bubbles, and molecular mass
It was observed that, over a long duration, the activity transport across the solvent-water interface. These
in the stagnant solvent phase becomes greater than that mechanisms are illustrated in Figure 2a and discussed
in the water phase, in which case, the driving force for below.
mass transport is reversed. It was recommended that Transport by Air Bubbles. The most significant
future studies should investigate ways to produce and pathway for the removal of pollutants is the unidirec-
maintain small bubbles at high air flow rates and reduce tional transport by air bubbles. As air bubbles rise up
backmixing from the organic solvent to the aqueous the column, both volatile and nonvolatile compounds
phase by continuously replenishing the solvent phase. partition into and on the surface of the bubble. For
These results have motivated our group to further dilute systems of volatile compounds, Henry’s law
investigate aspects of solvent sublation at the pilot-scale. adequately describes the extent of partitioning and,
These aspects include (i) the development of a steady- ultimately, the extent of stripping. However, for the
state mathematical model more suited for engineering case of strongly hydrophobic compounds that are of low
applications and scaleup; (ii) the effect of novel bubble volatility such as polyaromatic hydrocarbons, chlori-
spargers on the bubble size and other hydrodynamic nated pesticides, etc., the partitioning process is pri-
properties, and (iii) the effect of air, solvent, and water marily due to interfacial adsorption upon the bubble
flow rates on the performance of a three-phase continu- surface. The extent of partitioning can be determined
ous sublation column. The compounds chosen to be by combining the equilibrium relationships for bulk-
investigated in this research are pyrene and penta- phase partitioning and surface adsorption with a mass
chlorophenol. These pollutants have unique physio- balance for the total amount of pollutant carried by a
chemical properties as shown in Table 2. bubble of radius a. In eq 1, m is the total amount of
In part 1 of this series, we discuss the mechanisms pollutant, Γ is the surface concentration of the bubble,
of transport in solvent sublation and develop math- and Cv is the concentration of pollutant in the air
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1691


results:

m
CA )
4 3
πa
) (a3K
A )
+ Hc Cw (3)
3
The term in parentheses is the effective Henry’s law
constant, H. It depends inversely upon the air bubble
size. The ratio of the Henry’s law constants, H/Hc, is
termed the enhancement factor, H. From the values
of Hc and KA shown in Table 2 for pyrene, one can
determine that H is at least an order of magnitude
greater than Hc. When the bubble radius is on the order
of 0.05 cm, H becomes over a 100 times greater than
Hc.
Bubble-Wake Entrainment. Bubble-wake en-
trainment is the second most important transport
mechanism in solvent sublation. It results from the fact
that a bubble rising in a water column is not exactly
spherical. It is actually shaped more like a semispheri-
cal cap. Located behind the bubble resides a small
quantity of water called the wake (Fan and Tsuchiya,
1990), the amount of which is defined by the streamlines
in the fluid created by the rising bubble. When the
bubble and wake enter the solvent, the wake disengages
and becomes a small droplet that slowly descends to the
aqueous phase. This releases pollutant back into the
solvent. The droplet returns to the aqueous phase with
an equilibrium amount of pollutant. To account for this
type of transport, it is assumed that the wake uniformly
surrounds the bubble such that the volume of the wake
can be approximated as 4πa2di, where di is the wake
thickness.
Solvent-Water Interfacial Molecular Mass
Transfer. Unarguably, the organic solvent-water
interface is the focal point for all mass transport. As
shown in Figure 2a, one of the transport mechanisms
across the interface is molecular mass transport. In a
conventional liquid-liquid extraction process, this is the
sole mechanism of transport and is very important.
However, it is not as important in solvent sublation. The
explanation for this is that, in a sublation column,
convective mass transport due to the rising air bubbles
overwhelms the rate of molecular mass transport and
prevents the establishment of equilibrium across the
solvent-water interface. This is only true if the sub-
lation column is operated in a three-phase continuous
mode where steady-state operation is possible.

Model Development

The traditional (Lionel et al., 1981; Womack et al.,


1982; Wilson and Valsaraj, 1982-3; Huang et al., 1983;
Figure 2. (a) Mechanisms of mass transport, (b) series CSTR Valsaraj and Thibodeaux, 1991a,b) mathematical analy-
model (SCM), and (c) two-phase axial dispersion model (ADM2). sis of sublation columns involves dividing the column
into a series of well-mixed stages, the number of which
4 describes the degree of axial mixing, and simultaneously
m ) 4πa2Γ + πa3Cv (1) solving a set of mass balance equations. One non-
3
traditional approach reported in the literature
bubble. At equilibrium, the concentrations are related (Stachurski and Szeglowski, 1974) involves probabilistic
to the bulk-phase water concentration through linear arguments to predict removal efficiencies. Though there
relationships is nothing inherently wrong with either approach, they
are not generally accepted methods of analyzing gas-
Γ ) KACw Cv ) HcCw (2) liquid mass transfer in bubble columns. The preferred
approach is to treat the water column as a continuum
where KA is the interfacial partition constant, and Hc and to use the axial dispersion model (ADM) (Deckwer
is the Henry’s law constant. When eqs 1 and 2 are et al., 1983; Deckwer and Schumpe, 1993). In the
combined, the following effective air concentration following development, we apply both the traditional
+ +

1692 Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996

model and the axial dispersion model to solvent subla- persion coefficient and a second derivative. In this
tion. We further show that one can easily convert from development, axial dispersion in the gas phase is also
one model to the other. included.
Series CSTR Model (SCM). A general illustration Figure 2c shows a depiction similar to that used in
of the SCM appears in Figure 2b. The top stage, the SCM; however, the transport mechanisms within
designated as stage 1, represents the thin layer of the column have been drawn about a differential ele-
solvent which floats upon the column of water. The ment, dx. The solvent is not considered part of the
stages below the first make up the water column, which continuum and is treated as a separate well-mixed box.
is partitioned into an air phase and a water phase. The The governing equations that describe the transport of
volume fraction of the air phase is equal to the gas pollutant are
holdup, , while that of the water phase is 1 - . The
total number of stages comprising the water column ∂A ∂2A ∂A
determines the amount of axial mixing present in the ) 2 + Pe1 + Sh1(W - A) (10)
∂τ ∂ξ ∂ξ
column. The simplest case to consider is a two-staged
model (or two-box model); i.e., the water column is
comprised of just one stage (box). ∂W ∂2W ∂W
) 2 - Pe2 - Sh2(W - A) (11)
The transport mechanisms shown in Figure 2b in- ∂τ ∂ξ ∂ξ
clude convection, molecular-mass transfer between the
water and solvent, molecular-mass transfer between the 1 dS
) PA(HA|ξ)0 - S) + (Eo + Sto)(W|ξ)0 - S) - S
water and air, and bubble-wake entrainment. These φ dτ
mechanisms are included in steady-state mass balances (12)
that describe the transport of pollutant in the sublation
Equations 10 and 11 apply to the air and water phases
process. For a two-box SCM, the governing mass
and are referred to here as the two-phase axial disper-
balances are
sion model (ADM2). Equation 12 is a mass balance

( )
around the solvent layer. When the time derivatives
3k 2 CA in eqs 10-12 are set to zero and the appropriate
QwCw° - QwCw - πrc L Cw -
a H
-
boundary conditions are used, a steady-state solution

( )( )
di Co can be found as shown in the appendix. In the general
πrc2k1(1 - ) + 3QA Cw - ) 0 (4) case, the Peclet and Sherwood numbers would be
a Kow written in terms of two different axial dispersion coef-
ficients, one for the water phase and one for the air

QACA° - QACA +
3k 2
a (
πrc L Cw -
CA
H
)0) (5)
phase. Since it will be shown in part 2 of this series
that the bubbly flow regime is the preferred regime for
solvent sublation, the gas-phase dispersion coefficient

( )
is assumed to be equal to that of the liquid phase. This
HcCo is a good assumption for the bubbly flow regime;
QoCo° - QoCo + QA CA -
Kow
+
however, it has been reported (Shetty et al., 1992) that

( )( )
di Co the mechanism of gas-phase dispersion in the hetero-
πrc2k1(1 - ) + 3QA Cw - ) 0 (6) geneous regime (characterized by two distinct bubble
a Kow sizes) is distinctly different from that in the homoge-
neous regime. Therefore, the single-dispersion-coef-
where Cw, CA, and Co are the effluent water, air, and ficient assumption is not valid in the heterogeneous
solvent molar concentrations, respectively, and Qw, QA, regime.
and Qo are the water, air, and solvent volumetric flow A relationship exists (Fogler, 1992) that allows one
rates, respectively. The superscript degree, designates to determine the axial dispersion coefficient used in the
the molar feed concentration; however, in this analysis, ADM2 from the number of well-mixed boxes used in the
it is assumed that the air and solvent enter the process SCM. The derivation of the relationship involves de-
free of pollutant (i.e., CA° ) Co° ) 0). When eqs 4-6 termining the effluent response curves (or residence
are divided through by QwCw°, QAHCw°, and QoKowCw°, time distributions, RTDs) of both models when the
respectively, the following dimensionless equations mass-transfer terms are set to zero, and the feeds to
result: each model are subjected to a perfect impulse. Next,
the variance (second moment) of the RTD for n-CSTRs
1 - W - StwA(W - A) - (Stw + Ew)(W - S) ) 0 (7) in series is set equal to the variance of the RTD for the
axial dispersion model. The result is eq 13. After
A° - A + StA(W - A) ) 0 (8) 1 2 8
(13)
n Pe1 Pe 2
) +
1
S° - S + (Sto + Eo)(W - S) + PA(HA - S) ) 0 (9)
rearrangement, the expression for the dispersion coef-
The system of eqs 7-9 has an analytical solution which ficient is obtained as
is given in the appendix. When the number of boxes
used in the SCM is greater than two, a numerical uL
D) (14)
algorithm is recommended. n + (n2 + 8n)1/2
Two-Phase Axial Dispersion Model (ADM2). An
alternative method for modeling the sublation process where n is the number of CSTRs making up the water
is to consider the water column as a continuum of well- column (n ) i - 1), u is the gas velocity (ug/), and L is
mixed, interacting boxes. The interactions allow for the length of the vessel. The obvious benefit of this
axial mixing which is defined in terms of a dis- relationship is that one can easily convert between the
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1693

Figure 3. Fractional removal versus QA and Qw as predicted by Figure 4. Fractional removal versus QA/Qw shows multiple
the SCM. steady-states.
Table 3. Model Parameters Geary and Rice, 1991). The bubble-wake correlation
thermodynamic kinetic hydrodynamic design appearing in the table possesses the same functional
Hc ) 5 × 10-4 k ) 5 × 10-4 cm/s a ) 0.1ug1/2 cm rc ) 5.08 cm form of experimental data cited elsewhere (Fan and
KA ) 5 × 10-3 cm k1 ) 10-4 cm/s  ) 0.05ug L ) 152.4 cm Tsuchiya, 1990). Overall, the characteristics of the
Kow ) 76412 n)1 Vw/Vb ) 1.5 × bubbly flow surface plot are what one would expect: fractional
10-4Vb-0.741 type sparger removal varies directly with air flow rate and inversely
with water flow rate; and the approach toward unity is
two models given either the number of CSTRs or the asymptotic. Though the surface plot nicely illustrates
axial dispersion coefficient. the model, it is not accessible for quantitative use.
Furthermore, multiple surface plots are required to span
Results and Discussion Qo. Because of these reasons, an alternate way to
present the model is necessary.
The two response functions that describe the ef- Since it has been customary in previous sublation
fectiveness or performance of a sublation column are studies to plot fractional removal versus a ratio of flow
the fractional removal of pollutant, FR, and the separa- rates, the same fractional removal data used in the
tion factor, Σ, previous figure was plotted versus QA/Qw (see Figure
4). As can be seen in the figure, multiple steady states
Cw are predicted. The reason why this occurs is that
FR ) 1 - W ) 1 -
Cw° fractional removal is affected by column hydrodynamic
properties which are functions of gas velocitysnot QA/
S Co Qw. For example, the hydrodynamics of the column are
Σ) (15) different at (QA ) 2000 cm3/min; Qw ) 20 cm3/min) than
W CwKow
)
they are at (QA ) 10000 cm3/min; Qw ) 100 cm3/min)
even though the ratios of QA/Qw are the same. If Qw
The separation factor is the ratio of the effluent solvent were fixed, then multiple steady states would not occur
concentration to the effluent water concentration and as the ratio QA/Qw would be directly proportional to gas
is a good measure of the effectiveness of solvent subla- velocity. Again, the problem with fixing a variable is
tion relative to other liquid-liquid separation processes. that multiple plots become necessary. Therefore, it was
Both response functions are dependent upon five types decided to correlate the model predictions in terms of
of variables: thermodynamic, kinetic, design, opera- the dimensionless numbers that describe the mass
tional, and hydrodynamic. Not all of these variables, transport mechanisms. The advantage of this approach
however, are independent. In most cases, the kinetic is that the effect of the operational variables and the
and hydrodynamic variables are defined in terms of the effect of their influences on column hydrodynamics are
gas velocity, which is an operational variable. Hydro- combined. Based on multiple-variable regression analy-
dynamic properties are also dependent upon the design ses of 250 computer simulations, correlations for frac-
variables, especially the sparger type. Thus, the prob- tional removal and the separation factor were deter-
lem of characterizing sublation column performance is mined. The ranges of the operational variables used
one which investigates the effects of the operational, in the simulations were 400 e QA e 15 200 cm3/min,
design, and thermodynamic variables. 20 e Qw e 100 cm3/min, and 0.1 e Qo e 100 cm3/min.
Effect of Operational Variables (QA, Qw, Qo). In For strongly hydrophobic compounds where the solute
Figure 3, a representative surface of fractional removal capacity of the solvent is large (Pw e 0.013), a simple
versus QA and Qw is shown for a strongly hydrophobic correlation for the fractional removal was found.
compound, like a polynuclear aromatic hydrocarbon
(PAH). The surface was generated from the SCM using FR
typical values (i ) 2 and Qo ) 2 cm3/min). An equiva- ) Ew0.1657StwA0.8352 (16)
1 - FR
lent surface would result from the ADM2 if Pe1 ) 4.
Other model parameters that were used to generate the In Figure 5a, a parity plot of fractional removal shows
surface are listed in Table 3. As shown in the table, very good agreement between the theory and the
the hydrodynamic variables were determined from correlation, as all the deviations are within (10%. The
power-law models which are representative of the reason for the good agreement is that the functional
bubbly flow regime (Clift et al., 1978; Shah et al., 1982; dependence of eq 16 is consistent with the relative
+ +

1694 Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996

the terms, leaving a net (QoKow)-1 dependence in the


numerator and denominator.

ΨKow-1Qo-1
Σ∼ (19)
1 + ΨKow-1Qo-1

where

[ ]QAdi 1.464
Ψ≡ 3 [QAHc]0.544 (19)
a
The physical importance of eq 19 becomes clear when
one recognizes that the terms remaining in the brackets
(collectively referred to as the kinetic factor, Ψ) reflect
the kinetic and hydrodynamic processes associated with
the air (bubble) phase. Essentially, Ψ controls the
extent to which equilibrium is approached between the
solvent and water. First, in the limit of small Ψ, the
departure from equilibrium is great, and as eq 19
suggests, Σ is predicted to be zero. Of course, a zero
separation factor is not accurate since the physics of the
process change significantly when the air phase is
removed. The lower limit is better described by con-
sidering the solvent and water phases as two contacting
CSTRs. Though such a situation is kinetically undesir-
able in practice, it does represent a theoretical lower
limit for solvent sublation and does exist for a single-
staged quiescent solvent extraction process. It can be
Figure 5. Parity plots of the theoretical (a) fractional removal shown that, for such a case, Σ is equal to Sto/(1 + Sto).
and (b) separation factor versus predictions by correlation. Using this expression and typical conditions used in this
paper (Sto ) 6.34 × 10-6), one can estimate that the
importance of the mass-transfer mechanisms. The mass lower limit for Σ ≈ Sto , or equivlalently Σ/(1 - Σ) ≈ Sto.
transfer associated with the air bubbles, StwA, is weighted As one can see in Figure 5b, predictions of Σ for practical
more than that associated with entrainment, Ew. As sublation conditions are as much as 3 orders of magni-
alluded to, the mass transfer across the solvent-water tude greater than the lower theoretical limit. The
interface (Stw or Sto) has no effect on the fractional improved separation is attributed to the additional
removal for systems which are overwhelmed by the transport processes associated with the air (bubble)
transport of air bubbles. One advantage of the correla- phase. In the limit of large Ψ, the separation factor
tion is that fractional removal of strongly hydrophobic approaches unity, the condition for solvent-water equi-
compounds can be estimated very easily without resort- librium. In practice, however, it is unlikely that equi-
ing to the analytic solution. Another advantage is that librium may be reached solely by adjusting Ψ. Since
the correlation suggests the appropriate way in which the hydrodynamic properties were assumed to be that
experimental data should be plotted, as will be shown of homogeneous flow, the SCM will eventually break
in a later paper (part 3 of this series). down as Ψ becomes very large. Thus, other means of
From the regression analysis, a simple correlation for improving the performance are needed, such as reducing
the separation factor was also obtained. the solvent usage, Qo, without significantly affecting the
overall fractional removal.
Σ It is emphasized that the correlations, eqs 16 and 17,
) Eo1.464PA0.544QoKow (Qo in cm3/min) (17)
1-Σ were developed from the SCM with fixed kinetic pa-
rameters (see Table 3) and are valid for strongly
A parity plot of this correlation and the theoretical hydrophobic compounds (Pw e 0.013). For weakly
separation factor is shown in Figure 5b. Again, the hydrophobic compounds, the exact solutions given in the
agreement is within (10%. One interesting feature appendix must be used.
about the correlation is that, though it is appears first Effect of Thermodynamic Variables. Examining
order in QoKow, the overall dependence is actually the effects of thermodynamic properties is important
≈(QoKow)-1. To explicitly see the inverse dependence, because they can vary greatly between families of
eq 17 is rearranged and expressed in terms of dimen- hydrophobic pollutants. As shown in Table 2, the
sional numbers. interfacial partition constant can vary over 4 orders of
magnitude. Other properties such as the Henry’s law

Co
3 {
QoKowa } { }
QAdi 1.464 QAHc 0.544
QoKow
QoKow
constant and the solvent-water partition coefficient can
vary greatly as well. Therefore, computer simulations,

{ } { }
Σ) again based on the SCM (i ) 2) and values cited in Table
CwKow QAdi 1.464 QAHc 0.544
)
3, were performed to determine the relative importance
1+ 3 QoKow of these thermodynamic properties.
QoKowa QoKow
(18) There was negligible effect of the solvent-water
partition constant on fractional removal for Kow g 700.
Observing that the exponents for the terms in braces This was not a surprise given the fact that fractional
sum to approximately 2, one can factor a (QoKow)-2 from removal was shown not to be dependent upon solvent-
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1695

Figure 7. Effect of bubble size on fractional removal in solvent


sublation.

One interesting feature about Figure 6 is that each


QA/Qw curve represents not only a fixed air flow rate
but also a fixed bubble size, since the bubble size was
taken to be a function of gas velocity. The implication
is that one can predict the effect of a smaller bubble
size at a constant air flow rate or, similarly, the effect
of increasing the air flow rate at constant bubble size.
Such an exercise may be important if surfactant is used
to change the surface tension or if an improvement in
sparger design is made, as will be discussed in the next
section. To illustrate the idea, consider curve (b) in
Figure 6b at H ) 10. The fractional removal is about
80% for these conditions. The bubble size can be shown
to be about 1 mm. If the bubble size were to be made
50% smaller, H would become 19, corresponding to a
Figure 6. Effect of the Henry’s enhancement factor on fractional fractional removal of 86%. If the bubble size were an
removal (Qw ) 50 cm3/min, Qo ) 2 cm3/min) for (a) Hc ) 10-4, (b) order of magnitude smaller than its original value at
Hc ) 10-3, and (c) Hc ) 10-2. this flow rate, the fractional removal would increase to
water mass transfer for strongly hydrophobic com- 91%. This illustration demonstrates that bubble size
pounds. However, the effects of Hc and KA on fractional affects not only the interfacial area for mass transfer
removal were significant. In Figure 6a, fractional but also the thermodynamics (enhancement factor)
removal is plotted versus the enhancement factor, H ) influencing the driving force for mass transfer.
1 + 3KA/aHc, for constant values of QA/Qw. The en- Effect of Design Variables. Design variables do not
hancement factor was varied by varying KA over the play a large role in solvent sublation; however, they can
range 5 × 10-5-5 × 10-2 cm. As expected, the enhance- affect hydrodynamic properties which do have signifi-
ment factor had a pronounced favorable effect on cant impacts on sublation performance. For example,
fractional removal. Even for compounds which do not bubble size is determined, to a great extent, by the type
have large enhancement factors, the effect was ap- of device used for bubble generation. Since bubble size
preciable at high air flow rates. At low air flow rates, contributes largely to the interfacial area available for
the effect was significantly reduced even for H on the mass transfer as well as the enhancement factor, the
order of 100. The condition where H ) 1 represented choice of bubble generator is important. Some examples
the situation where no adsorption to the bubble occurs. of bubble generators are submersible orifices (gas sparg-
In this situation, the mass transfer to the air phase was ers) such as nozzles, sintered-metal or -glass plates,
solely driven by Henry’s law. sprinklers, and rubber membranes. Some bubble gen-
For the case shown in Figure 6a, Hc was 10-4. When erators are external to the column. An example is an
values higher than 10-4 were used, the curves in Figure aerated vessel containing surfactant and water. The
6a were simply translated up the x-axis by an amount bubbles which form in the vessel are removed and
of 1/∆Hc. For example, when Hc was 10-3, the curves injected into the bottom of the bubble column.
were translated up the x-axis by a factor of 10, as shown Small improvements in sparger design may have
in Figure 6b. The shape of the curves and the values significant effects on sublation performance. An ex-
of fractional removal remained unchanged. For Hc > ample of this is illustrated in Figure 7, where fractional
10-3, the curves were translated up the x-axis such that removal is plotted versus QA/Qw for two different power-
the lower portion of the curves were truncated. This is law models describing bubble size dependence: one with
illustrated in Figure 6c, where fractional removal is a square-root dependence on velocity and one with a 1/4
plotted for Hc ) 10-2. The reason for this behavior is power dependence. The model with the 1/4 dependence
that, for a small Hc, surface adsorption is the sole mass is hypothetical; however, it is not necessarily infeasible.
transfer mechanism; thus, to achieve a high degree of Spargers that eject bubbles into an overlying shear field
fractional removal, a large enhancement factor is re- have been shown to produce smaller bubbles than what
quired. For a large Hc, however, Henry law partitioning is normally observed in the bubbly flow regime (Johnson
becomes significant, and the same degree of fractional et al., 1982; Johnson and Gershey, 1990; Marshall et
removal is achieved at a lower enhancement factor. al., 1993). Such spargers may well be described by a
+ +

1696 Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996

Table 4. Comparison of Fractional Removal between the


SCM (i ) 2) and the ADM2 (Pe1 ) 4)
QA (cm3/min) Qw (cm3/min) Qo (cm3/min) FR ADM2 FR SCM
400 40 20 0.799 0.786
400 20 2 0.888 0.880
400 80 2 0.665 0.647
400 40 0.01 0.788 0.774
3200 80 2 0.809 0.791
3200 40 20 0.894 0.883
3200 20 2 0.944 0.938
3200 40 0.01 0.889 0.877
6200 20 2 0.957 0.952
6200 80 2 0.849 0.833
6200 40 0.01 0.914 0.905
6200 40 20 0.918 0.909
12200 80 2 0.884 0.871
12200 40 20 0.939 0.931
Figure 8. Effect of axial mixing (Pe1) on fractional removal. 12200 40 0.01 0.936 0.928
12200 20 2 0.968 0.964
power-law model with a weaker dependence in velocity.
In part 2 of this series, bubble size results from an boundaries instead of specifying the concentration
annular shear bubble sparger support this type of model (closed vessel). When one examines an aqueous con-
dependence. centration profile generated from the ADM2 for the
Gas holdup is an important hydrodynamic property conditions given, a slight gradient is observed. Remov-
that contributes to the interfacial area. In general, it ing the gradient would bring the fractional removal
is determined soley by the gas velocity and the physical values sufficiently close together. The implication is
properties of the liquid (Akita and Yoshida, 1973; Hikita that the appropriate boundary conditions for the ADM2
et al., 1980; Shah et al., 1982). However, in part 2, be “closed” for the case where the model is compared to
experimental results indicate that gas holdup may be the SCM with n ) 1 (i.e., no convective character at all).
improved by the use of novel spargers, such as the Despite the small offset, the results in Table 4 indicate
annular shear sparger already mentioned. that the two mathematical models are indeed equivalent
Axial dispersion, on the other hand, has a strong and can be used interchangeably.
dependence upon column diameter (Deckwer and
Schumpe, 1993). In general, reducing the axial disper-
sion or mixing in bubble column reactors is desirable Conclusions
and can be accomplished using tall columns (Deckwer
et al., 1974). To elucidate the effect of axial mixing on Solvent sublation is a nonfoaming adsorptive bubble
sublation performance, the ADM2 was used to show the process which is capable of removing trace levels of
dependence of fractional removal on the Peclet number, nonvolatile and volatile organic compounds from waste-
Pe1. As one can see in Figure 8, there is essentially no waters. The advantage of solvent sublation over bubble
effect of mixing. This behavior is consistent with fractionation or air stripping is that higher removal
experimental observations (Valsaraj et al., 1992) which efficiencies are possible. Like liquid-liquid extraction,
showed that the fractional removal was independent of sublation utilizes a water-immiscible solvent for con-
column diameter. The explanation for this is that the taminant recovery; however, in sublation, mixers and
solute is strongly retained in the organic phase and is phase separators are not needed. Moreover, the effluent
prevented from being backmixed in the aqueous phase. water from a sublation column does not require further
Therefore, unlike other bubble column processes, the treatment to remove residual solvent.
influence of column diameter on axial mixing is not In this paper, the transport mechanisms in solvent
important. From a mathematical perspective, the im- sublation were presented. In order of relative impor-
plication is that the two-box SCM should adequately tance, the mechanisms are (i) transport by air bubbles,
describe sublation performance despite the extent of (ii) transport by bubble-wake entrainment, and (iii)
mixing. In part 3 of this series, experimental data transport by solvent-water mass transfer. These mech-
obtained from a pilot sublation process compare well anisms were combined with mass balances derived from
with the SCM and support the idea that mixing in the a continuous, countercurrent sublation process. Two
aqueous phase is unimportant in sublation processes. equivalent, mathematical models resulted, namely, the
ADM2 versus SCM. To illustrate how well the series CSTR model (SCM) and two-phase axial disper-
ADM2 compares with the SCM, fractional removal data sion model (ADM2). Based upon hydrodynamic models
were calculated from each model using the analytical characterizing the bubbly flow regime in bubble col-
solutions listed in the appendix. The calculations were umns, the SCM was used to elucidate the effects of
based on the model parameters listed in Table 3. The operational, thermodynamic, and design variables on
axial dispersion coefficient was determined from eq 14. solvent sublation performance. It was found that these
Table 4 shows the various conditions for the operating effects can be approximated in terms of correlations
variables and the corresponding model results. As can involving dimensionless numbers that describe the
be seen in the table, the ADM2 consistently predicts a transport mechanisms.
slightly higher fractional removal than the SCM but is For strongly hydrophobic compounds (Pw e 0.013), the
no higher than about 0.01 in all but one case. One steady-state fractional removal was easily determined
possible explanation for this offset may be attributed from the air-water Stanton number, StwA, and the
to the choice of boundary conditions used for the ADM2. water-side entrainment number, Ew. Determining the
The traditional Danckwerts conditions (open vessel) separation factor required the oil-side entrainment
were used in the development. These conditions require number, Eo, the air-solvent capacity factor, PA, and the
that the dispersive flux match the convective flux at the product QoKow (not dimensionless).
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1697


The effect of thermodynamic properties on sublation StwA ) water-side Stanton number (air-water), StwA ≡
performance was illustrated by plotting the fractional 3kπrc2L/aQw
removal versus the enhancement factor, H, which te ) exposure time
accounts for the surface-active nature of compounds. It ug ) superficial gas velocity, cm/s
was shown that large enhancement factors have favor-
Vb ) bubble volume, cm3
able effects on fractional removal for nonvolatile com-
pounds. However, as the volatility increases, the effect Vo ) volume of solvent, cm3
of the enhancement factor becomes less important. Vw ) wake volume, cm3
Finally, the effects of design variables were discussed. W ) dimensionless water concentration, W ≡ Cw/Cw°
It was shown that small improvements in sparger Greek Letters
design (i.e., producing smaller bubbles) can have a
significant, favorable effect on sublation performance.  ) gas holdup, cm3/cm3
Additionally, it was shown that the effect of the column H ) Henry’s enhancement factor, H ≡ 1 + 3KA/aHc
diameter on axial mixing does not influence sublation φ ) scale factor, QoL2/DVo
performance.
Γ ) surface concentration, mol/cm2
λ ) dispersivity, cm
Acknowledgment µ ) fluid viscosity, g/cm s
F ) fluid density, g/cm3
This work was supported in part by the National
σ ) surface tension, dyn/cm
Science Foundation/EPSCOR program {NSF/LaSER
(1992-96) ADP-03}. Σ ) separation factor, Σ ≡ Co/KowCw
τ ) dimensionless time, τ ≡ tD/L2
ξ ) dimensionless distance, ξ ≡ x/L
Nomenclature
Ψ ) kinetic factor
a ) bubble radius, cm
A ) dimensionless air concentration, A ≡ CA/HCw°
CA ) effective concentration of pollutant in air, mol/cm3
Appendix
Co ) concentration of pollutant in solvent, mol/cm3
Cv ) concentration of pollutant in air, mol/cm3 The analytic solution to the SCM with i ) 2 (n ) 1)
Cw ) concentration of pollutant in water, mol/cm3 is
CP ) chlorinated phenol
di ) bubble-wake thickness, cm
D ) dispersion coefficient, cm2/s W ) [(1 + StA)(1 + Sto + Eo + PA)]/
Eo ) solvent-side entrainment number, Eo ≡ 3QAdi/QoKowa [(1 + StA){(1 + PA)(1 + Ew + Stw) + Eo + Sto} +
Ew ) water-side entrainment number, Ew ≡ 3(QAdi/Qwa)
FR ) fractional removal, FR ≡ 1 - Cw/Cw° StwA(1 + Eo + Sto + PA) - StAPAH(Ew + Stw)]
H ) effective Henry’s law constant, Hc + 3KA/a
Hc ) Henry’s law constant, cm3/cm3
i ) number of stages comprising the SCM A ) [StA(1 + Sto + Eo + PA)]/
k ) overall mass-transfer coefficient (air-water), cm/s [(1 + StA){(1 + PA)(1 + Ew + Stw) + Eo + Sto} +
kl ) overall water-side mass-transfer coefficient (solvent-
water), cm/s StwA(1 + Eo + Sto + PA) - StAPAH(Ew + Stw)]
KA ) surface adsorption constant, cm
Kow ) solvent-water partition coefficient, cm3/cm3
Kow′′ ) octanol-water partition coefficient, cm3/cm3 S ) [(1 + StA)(Sto + Eo) + PAStAH]/
L ) length of bubble column, cm [(1 + StA){(1 + PA)(1 + Ew + Stw) + Eo + Sto} +
L/Dc ) column length to diameter ratio
m ) mass of pollutant carried by an air bubble, g StwA(1 + Eo + Sto + PA) - StAPAH(Ew + Stw)]
n ) number of stages comprising the water column,
n)i-1 By manipulating eq 15, one can express the fractional
PA ) air-solvent capacity factor, PA ≡ QAHc/QoKow removal and the separation factor in simple terms of
Pw ) water-solvent capacity factor, Pw ≡ Eo/Ew ≡ Qw/QoKow the water and solvent concentrations,
PAH ) polynuclear aromatic hydrocarbon
Pe1 ) air-phase Peclet number, Pe1 ≡ QAL/πrc2D
Pe2 ) water-phase Peclet number, Pe2 ≡ QwL/πrc2(1 - )D FR 1 Σ 1
QA ) air flow rate, cm3/s (cm3/min) ) -1
1 - FR W 1-Σ W
)
Qo ) solvent flow rate, cm3/s (cm3/min) -1
S
Qw ) water flow rate, cm3/s (cm3/min)
rc ) radius of bubble column, cm
S ) dimensionless solvent concentration, S ≡ Co/KowCw° When the analytic solution is substituted for W and S,
Sh1 ) air-phase Sherwood number, Sh1 ≡ 3kL2/aDH the exact solutions for the fractional removal and the
Sh2 ) water-phase Sherwood number, Sh2 ≡ (3kL2/aD)(/ separation factor result.
(1 - ))
StA ) air-side Stanton number (air-water), StA ≡ 3kπrc2L/
FR StwA
aQAH
Sto ) solvent-side Stanton number (solvent-water), Sto ≡ 1 - FR 1 + StA
) +

[ ][ ( )]
kπrc2(1 - )/QoKow Ew + Stw StA
Stw ) water-side Stanton number (solvent-water), Stw ≡ 1 + PA 1 - H
klπrc2(1 - )/Qw 1 + PA + Sto + Eo 1 + StA
+ +

1698 Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996

Eo + Sto + PAH
StA
[( ) ]
A ) λ4 1 +
λ4
Pe1
- eλ4ξ -

[( ) ] [( ) ]/
Σ 1 + StA

( )
λ2 λ3
1-Σ StA κ 2 eλ2 1 + - eλ2ξ + κ3 eλ3 1 + - eλ3ξ
)
1 + PA 1 - H Pe1 Pe1

[ ]{[ ( ) ]
1 + StA
γ1 - γ2 λ4 β4(γ1 - λ4) - γ2
eλ4 1 +
Pe2 Pe1
- -
γ1 - γ2

[( ) ]
The appropriate boundary conditions for the ADM2
are λ2 β2(γ1 - λ2) - γ2
κ2 eλ2 1 +
Pe1
- +
γ1 - γ2
dA
|
dξ ξ)0
)0
dW
|
dξ ξ)1
)0 [( ) κ3 λ3 1 +
λ3
Pe1
-
β3(γ1 - λ3) - γ2
γ1 - γ2 ]}
dA
| ) -Pe1Aξ)1 [( ) ]
W ) eλ4 1 +
λ4
Pe1
- β4eλ4ξ -

[( ) ] [( ]/
dξ ξ)1
κ2 eλ2 1 +
λ2
- β2eλ2ξ + κ3 eλ3 1 +
λ3
)
- β3eλ3ξ

|
Pe1 Pe1

[ ]{[ ( ) ]
dW
) Pe2(Ew + Stw)(W|ξ)0 - S) - Pe2(1 - W|ξ)0) γ1 - γ2 λ4 β4(γ1 - λ4) - γ2
dξ ξ)0 eλ4 1 +
Pe2 Pe1
- -
γ1 - γ2
where it has been assumed that the solvent and air
enter the column free of pollutant. The analytic solution [( )
κ2 eλ2 1 +
λ2
Pe1
-
β2(γ1 - λ2) - γ2
γ1 - γ2
+ ]
is given in terms of the following variables:

Pe1 1 2
[( ) κ3 eλ3 1 +
λ3
Pe1
-
β3(γ1 - λ3) - γ2
γ1 - γ2 ]}
β2 ) 1 - λ - λ where
Sh1 2 Sh1 2

x ( )( ) ( )( )
a1 λ4 β4eλ4 - β3eλ3 λ4 β4eλ4 - β2eλ2
λ2 ) 2 -Q cos (3θ + 240°) - 3 κ2 )
λ2 β eλ2 - β eλ3
κ3 )
λ3 β eλ2 - β eλ3
2 3 2 3

Pe1 1 2
β3 ) 1 - λ - λ and
Sh1 3 Sh1 3

x
a1 HPAA|ξ)0 + (Eo + Sto)W|ξ)0
λ3 ) 2 -Q cos (3θ + 120°) - 3 S)
1 + PA + Eo + Sto

x
Pe1 a1 Literature Cited
1 2
β4 ) 1 - λ - λ
Sh1 4 Sh1 4
λ4 ) 2 -Q cos (3θ) - 3 Akita, K.; Yoshida, F. Gas Holdup and Volumetric Mass Transfer
Coefficient in bubble Columns. Ind. Eng. Chem. Process Des.
Dev. 1973, 12, 76.
a1 ) Pe1 - Pe2 Caballero, M.; Cela, R.; Perez-Bustamante, J. A. Solvent Sublation
of Some Priority Pollutants (Phenols). Anal. Lett. 1988, 21 (1),
63-76.
a2 ) -(Pe1Pe2 + Sh1 + Sh2) Caballero, M.; Cela, R.; Perez-Bustamente, J. A. Studies on the
Use of Colloidal Gas Aphrons in Coflotation and Solvent
Sublation Processes. A Comparison with the Conventional
a3 ) Sh1Pe2 - Sh2Pe1 Technique. Sep. Sci. Technol. 1989, 24, 629-640.
Cervera, J.; Cela, R.; Perez-Bustamante, J. A. Analytical Solvent
Sublation of Metallic Dithizonates Part I. Solvent Sublation of
cos θ ) R/x-Q3 Copper. Analyst 1982, 107, 1425-1430.
Clift, R.; Grace, J. R.; Weber, M. E. Formation and Breakup of
Fluid Particles. Bubbles, Drops, and Particles; Academic
R ) (9a1a2 - 27a3 - 2a13)/54 Press: San Diego, 1978; pp 321-328.
Deckwer, W. D.; Schumpe, A. Improved Tools for Bubble Column
Reactor Design and Scale-Up. Chem. Eng. Sci. 1993, 48 (5),
889-911.
Q ) (3a2 - a12)/9 Deckwer, W. D.; Burckhart, R.; Zoll, G. Mixing and Mass Transfer
in Tall Bubble Columns. Chem. Eng. Sci. 1974, 29, 2177-2188.

[ ( )]
Deckwer, W. D.; Tien, K. N.; Kelkar, B. G.; Shah, Y. T. Applicabil-
Eo + Sto ity of Axial Dispersion Model to Analyze Mass Transfer Meas-
γ1 ) Pe2 1 + (Ew + Stw) 1 - urements in Bubble Columns. AIChE J. 1983, 29 (6), 915-921.
1 + PA + Eo + Sto
Fan, L. S.; Tsuchiya, K. Wake Sizes. In Bubble Wake Dynamics in
Liquids and Liquid-Solid Suspensions; Brenner, H., Ed.;
PAH Butterworth-Heinemann: Stoneham, MA, 1990; Chapter 5.
γ2 ) Pe2(Ew + Stw) Finch, J. A.; Dobby, G. S. Column Flotation; Pergamon Press:
1 + PA + Eo + Sto Toronto, 1990; Chapter 1.
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1699


Fogler, H. S. Analysis of Nonideal Reactors. Elements of Chemical Shetty, S. A.; Kantak, M. V.; Kelkar, B. G. Gas-Phase Backmixing
Reaction Engineering; Prentice-Hall: Englewood Cliffs, NJ, in Bubble-Column Reactors. AIChE J. 1992, 38 (7), 1013-1026.
1992; Chapter 14. Sheu, G. L.; Huang, S. D. Solvent Sublation and Adsorbing Colloid
Foltz, L. K.; Carter, K. N.; Wilson, D. J. Removal of Refractory Flotation of Magenta. Sep. Sci. Technol. 1987, 22 (11), 2253-
Organics by Aeration. VII. Solvent Sublation of Indene and 2262.
Aldrin. Sep. Sci. Technol. 1986, 21 (1), 57-78. Shih, K. Y.; Han, W. D.; Huang, S. D. Solvent Sublation of
Geary, N. W.; Rice, R. G. Bubble Size Prediction for Rigid and Hexachlorobutadiene and 2,4,6-Trichlorophenol. Sep. Sci. Tech.
Flexible Spargers. AIChE J. 1991, 37 (2), 161-168. 1990, 25 (4), 477-487.
Hikita, J.; Asai, K.; Tanigawa, K.; Segawa, K.; Kitao, M. Gas Shin, H. S.; Coughlin, R. W. Removal of Organic Compounds from
Holdup in Bubble Column. J. Chem. Eng. 1980, 20, 59. Water by Solvent Sublation. J. Colloid Interface Sci. 1990, 138
Hoff, J. T.; Mackay, D.; Gillham, R.; Shiu, W. Y. Partitioning of (1), 105-112.
Organic Chemicals at the Air-Water Interface in Environmental
Spargo, P. E.; Pinfold, T. A. Studies in the Mechanism of Sublate
Systems. Environ. Sci. Technol. 1993, 27 (10), 2174-2180.
Removal by Solvent Sublation. Part II. Sep. Sci. 1970, 5 (5),
Huang, S. D.; Valsaraj, K. T.; Wilson, D. J. Removal of Refractory
619-635.
Organics by Aeration. V. Solvent Sublation of Naphthalene and
Phenanthrene. Sep. Sci. Technol. 1983, 18 (10), 941-968. Stachurski, J.; Szeglowski, Z. Verification of a Stochastic Model
Huiru, D.; Xiuyu, Y. Solvent SublationsSpectrophotometric De- for a Frothless Solvent Ion Flotation Using Thulium and
termination of Zinc(II) with Thiocyanate and Malachite Green. Americium. Sep. Sci. 1974, 9 (4), 313-324.
Anal. Lett. 1988, 21 (6), 1065-1073. Tamamushi, K.; Wilson, D. J. Removal of Refractory Organics by
Johnson, B. D.; Gershey, R. M. Bubble Formation at a Cylindrical Aeration. VI. Solvent Sublation of Alkyl Phthalates. Sep. Sci.
Frit Surface in a Shear Field. Chem. Eng. Sci. 1990, 46 (10), Technol. 1984-5, 19, 1013-1023.
2753-2756. Valsaraj, K. T. Removal of Organics from Water by Non-foaming
Johnson, B. D.; Gershey, R. M.; Cooke, R. C.; Sutcliffe, W. H. A Flotation In Flotation-Science and Engineering; Matis, K. A.,
Theoretical Model for Bubble Formation at a Frit Surface in a Ed.; Marcel Dekker: New York, 1995; pp 365-383.
Shear Field. Sep. Sci. Technol. 1982, 17 (8), 1027-1039. Valsaraj, K. T.; Wilson, D. J. Removal of Refractory Organics by
Karger, B. L. In Adsorptive Bubble Separation Techniques; Lem- Aeration. IV. Solvent Sublation of Chlorinated Organics and
lich, R., Ed.; Academic Press: New York, 1972, Chapter 8. Nitrophenols. Colloids Surf. 1983, 8, 203.
Karger, B. L.; Caragay, A. B.; Lee, S. B. Studies in Solvent Valsaraj, K. T.; Springer, C. Removal of Traces of Pentachlorphenol
Sublation: Extraction of Methyl Orange and Rhodamine B. Sep. from Aqueous Acidic Solutions by Sovlent Extraction and
Sci. 1967a, 2 (1), 39-64. Sovlent Sublation. Sep. Sci. Technol. 1986, 21 (8), 789-807.
Karger, B. L.; Grieves, R. B.; Lemlich, R.; Rubin, A. J.; Sebba, F. Valsaraj, K. T.; Thibodeaux, L. J. Studies in Batch and Continuous
Nomenclature Recommendations for Adsorptive Bubble Separa- Solvent Sublation. I. A Complete Model and Mechanisms of
tion Methods. Sep. Sci. Technol. 1967b, 2, 401. Sublation of Neutral and Ionic Species from Aqueous Solutions.
Karger, B. L.; Pinfold, T. A.; Palmer, S. E. Studies in the Sep. Sci. Technol. 1991a, 26 (1), 37-58.
Mechanism of Sublate Removal by Solvent Sublation. Part I. Valsaraj, K. T.; Thibodeaux, L. J. Studies in Batch and Continuous
Sep. Sci. 1970, 5 (5), 603-617. Solvent Sublation. II. Continuous Countercurrent Solvent
Kown, B. T.; Wang, L. K. Solute Separation by Continuous Bubble Sublation of Neutral and Ionic Species from Aqueous Solutions.
Fractionation. Sep. Sci. 1971, 6 (4), 537-552. Sep. Sci. Technol. 1991b, 26 (3), 367-380.
Lemlich, R. A Theoretical Approach to Nonfoaming Adsorptive Valsaraj, K. T.; Thoma, G. J.; Thibodeaux, L. J.; Wilson, D. J.
Bubble Fractionation. AIChE J. 1966, 12 (4), 802-804. Nonfoaming Adsorptive Bubble Separation Processes. Sep. Tech-
Lionel, T.; Wilson, D. J.; Pearson, D. E. Removal of Refractroy nol. 1991, 1, 234-243.
Organics from Water by Aeration. I. Methyl Chloroform. Sep.
Valsaraj, K. T.; Lu, X. Y.; Thibodeaux, L. J. Continuous Nonfoam-
Sci. Technol. 1981, 16 (8), 907-935.
ing Adsorptive Bubble Separation Processes for the Removal of
Lu, X.; Valsaraj, K. T.; Thibodeaux, L. J. Studies in Batch and
Hydrophobic Organic Compounds from the Aqueous Phase:
Continuous Solvent Sublation. IV. Continuous Countercurrent
ACS Symp. Ser. 1992, 509, 116-128.
Solvent Sublation and Bubble Fractionation of Hydrophobic
Organics from Aqueous Solutions. Sep. Sci. Technol. 1991, 26 Wang, W. K.; Huang, S. D. Solvent Sublation and Adsorptive
(7), 977-989. Flotation/Sublation of Diphenyl. Sep. Sci. Technol. 1988, 23
Mackay, D.; Shiu, W.; Valsaraj, K. T.; Thibodeaux, L. J. Air-Water (4 & 5), 375-385.
Transfer: The Role of Partitioning. In Air Water Mass Transfer; Wang, Y.; Deng, T.; Ke, J. J. Gold Recovery with Solvent Sublation
Gulliver, J.; Wilhelm, S.; Eds.; American Society Civil from Chloride Solutions. Int. J. Miner. Proc. 1993, 40, 57-68.
Engineering: New York, 1990. Wilson, D. J.; Valsaraj, K. T. Removal of Refractory Organics by
Marshall, S. H.; Chudacek, M. W.; Bagster, D. F. A Model for Aeration. III. A Fast Algorithm for Modeling Solvent Sublation
Bubble Formation from an Orifice with Liquid Cross-Flow. Columns. Sep. Sci. Technol. 1982-3, 17 (12), 1387-1396.
Chem. Eng. Sci. 1993, 48 (11), 2049-2059. Womack, J. L.; Lichter, J. C.; Wilson, D. J. Removal of Refractory
Michelsen, D. L.; Ruettimann, K. W.; Hunter, K. R.; Sebba, F. Organics from Water by Aeration. II. Solvent Sublation of
Feasibility Study on the Use of Predispersed Solvent Extraction/ Methylene Blue and Methyl Orange. Sep. Sci. Technol. 1982,
Flotation Techniques for Removal of Organics from Wastewaters. 17 (7), 897-924.
Chem. Eng. Commun. 1986, 48, 155-163.
Montgomery, J. H.; Welkom, L. M. Groundwater Chemicals Desk Received for review June 16, 1995
Reference; Lewis Inc.: Chelsea, U.K., 1990; p 640. Accepted February 13, 1996X
Nolan, B. T.; McTernan, W. F. Application of Solvent Sublation
to the Simultaneous Removal of Emulsified Coal Tar and IE9503656
Dissolved Organics. Chem. Eng. Commun. 1988, 63, 1-15.
Sebba, F. Ion Flotation; Elsevier Publishing Co.: New York, 1962.
Shah, Y. T.; Kelkar, B. G.; Godbole, S. P.; Deckwer, W. D. Design
X Abstract published in Advance ACS Abstracts, April 1,
Parameter Estimations for Bubble Column Reactors. AIChE J.
1982, 28 (3), 353-379. 1996.

You might also like