Sei sulla pagina 1di 13

Journal of Membrane Science 361 (2010) 43–55

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Batch and continuous studies for ethyl lactate synthesis in a pervaporation


membrane reactor
Carla S.M. Pereira a , Viviana M.T.M. Silva a,∗ , Simão P. Pinho b , Alírio E. Rodrigues a
a
Laboratory of Separation and Reaction Engineering (LSRE), Associate Laboratory LSRE/LCM, Departamento de Engenharia Química,
Faculdade de Engenharia, Universidade do Porto, Rua Dr. Roberto Frias s/n, 4200-465 Porto, Portugal
b
Laboratory of Separation and Reaction Engineering (LSRE), Departamento de Tecnologia Química e Biológica,
Instituto Politécnico de Bragança, Campus de Santa Apolónia, 5301-857 Bragança, Portugal

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

Article history: Pervaporation process using commercial silica water selective membranes was evaluated to contribute
Received 20 November 2009 for the ethyl lactate process intensification by continuous pervaporation membrane reactor. Preliminary
Received in revised form 26 May 2010 studies were performed in order to assess the existence of membrane defects and mass transfer limita-
Accepted 10 June 2010
tions, studying the influence of feed pressure and flowrate, respectively. After, in the absence of mass
Available online 18 June 2010
transfer limitations, membrane performance was evaluated experimentally, at different composition
and temperature measuring the flux and selectivity of each species in binary mixtures (water/ethanol,
Keywords:
water/ethyl lactate and water/lactic acid). Thus, species permeances were obtained for each experiment
Ethyl lactate
Pervaporation membrane reactor
and correlated in order to account for the effect of temperature and feed composition. Permeances of
Concentration polarization ethanol and ethyl lactate depend solely on the temperature, following an Arrhenius equation; for water,
Temperature polarization its permeance follows a modified Arrhenius equation taking into account also the dependence on the feed
Non-isothermal model water content. Mathematical models, considering concentration and temperature polarization, and non-
Silica membrane isothermal effects as well, were developed and applied to analyze the performance of batch pervaporation
and continuous pervaporation membrane reactor, in both isothermal and non-isothermal conditions. The
PVMR with five membranes in series, operating at 70 ◦ C, leads to 98% of lactic acid conversion and 96%
of ethyl lactate purity.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction side. The partitioning and desorption steps are normally neglected
and the remaining steps are considered as the main contributions to
In the past few years, the interest in the application of reac- the overall resistance to mass transfer. The mass transfer resistance
tive separations to many chemical processes has substantially in membranes depends on its properties and also on the chem-
increased. This is particularly true for equilibrium limited reactions, ical and physical properties of the feed components. In the past,
where the removal of at least one of the reaction products shifts the the membrane materials and dimensions were not optimized and,
equilibrium towards the product formation. A variety of separation therefore, high membrane mass transfer resistance was commonly
techniques can be incorporated into the reactor; among all, perva- observed. However, new developments in membrane materials
poration is becoming a promising technology, potentially useful in and ultra-thin composite membranes have led to much smaller
applications such as dehydration of organic mixtures [1–4], sepa- membrane mass transfer resistances, which are now mainly due
ration of organic mixtures [5,6] and removal of organic compounds to the diffusive transport at the boundary layer, like noticed in
from aqueous solutions [7,8]. the removal of volatile organic compounds from wastewaters [9]
In pervaporation processes, the transport of the components and in the dehydration of cyclohexane [10]. Diffusive transport
from the feed liquid mixture to the vapor phase involves the follow- depends on the hydrodynamic conditions, solute and fluid physi-
ing steps: (i) mass transfer from the feed bulk to the feed membrane cal properties, and the system geometry. Several correlations based
interface; (ii) partition of penetrants between the feed and the on the Sherwood number for calculating the mass transfer coeffi-
membrane; (iii) selective transport (diffusion) through the mem- cient for transport in the boundary layer (Kbl ) have been proposed
brane; and (iv) desorption into the vapor phase on the permeate over the years for different membrane modules configurations
[11–18]. Pervaporation is not usually an isothermal process; and
temperature drop in the feed is frequently observed due to species
∗ Corresponding author. Tel.: +351 225081489; fax: +351 225081674. vaporization. The heat transfer in pervaporation involves: (i) heat
E-mail addresses: viviana.silva@fe.up.pt, vivsilva@fe.up.pt (V.M.T.M. Silva). transfer from the feed bulk to the feed membrane interface; (ii)

0376-7388/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2010.06.014
44 C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55

heat transfer through the membrane; and (iii) consumption of 2. Experimental


heat (vaporization followed by expansion) at the permeate side
of the membrane. In systems operating under laminar conditions, 2.1. Materials
where the membrane is thin and/or has high permeability, the
temperature drop in the feed membrane interface is considerable The chemicals used were ethanol (>99.9% in water), lactic acid
[19] and, therefore, the boundary layer heat transfer resistance (>85% in water) and ethyl lactate (>98% in water) from Sigma-
should be taken into account. The heat transfer coefficient in the Aldrich (U.K.). A commercial strong-acid ion-exchange resin named
boundary layer is often determined by Nusselt based correlations Amberlyst 15-wet (A15-wet) (Rohm & Haas) was used as catalyst
[20]. and adsorbent. A commercial hydrophilic membrane supplied by
Aiming to produce esters, where water is a by-product, suit- Pervatech BV (The Netherlands) was used. It is resistant for any sol-
able membranes for dehydration of organic mixtures can be applied vent at any concentration, but sensitive to acidic and alkaline media
for esters process intensification by means of pervaporation mem- (preferred pH range from 2 up to 9) and so, exposure to inorganic
brane reactors [21–27]. Although the significant number of works acids or caustics must be avoided. This membrane has a modified
dealing with PVMR in the last years, most of them use simplified silica (methyl silica) selective layer coated onto gamma alumina.
mathematical models to optimize and predict the behaviour of The separation layer is applied inside of an asymmetric ceramic
PVMR units, neglecting concentration and temperature polariza- tube that has an outer diameter of 10 mm, an inner diameter of
tion effects, which can be very significant under certain working 7 mm, and a length of 50 cm. It has an effective membrane area per
conditions [28]. tube of about 110 cm2 . Two equal membrane modules were used
The ethyl lactate synthesis on pervaporation and vapour- to perform the experiments.
permeation membrane reactors was studied before by some
authors. The two configurations adopted were: (i) batch reactor, 2.2. Pervaporation membrane unit
where the esterification reaction takes place, followed by a mem-
brane for water removal, and reflux of the retentate to the reactor The experiments were carried out in a pervaporation membrane
[29–32] and (ii) membrane inside a batch reactor [33,34]. In fact, unit which can operate in batch or in continuous mode. This unit
none of these studies considers a continuous integrated membrane is equipped with temperature (TI) (type K thermocouple, accuracy
reactor for ethyl lactate production. Regarding the type of mem- of about ±2.2 ◦ C) and pressure sensors (PI) in order to monitor and
branes tested, polymeric membranes were used for pervaporation register these two parameters. The absolute pressure is measured
[29,30,32] and zeolites for vapour-permeation [33,34]. However, through two analogue dials (accuracy of about ±0.5 bar), filled
generally, the pervaporation process is superficially addressed; with glycerine (Nuova Firma), while the pressure in the perme-
important membrane parameters, like selectivity and species per- ate side is measured by means of 1 digital dial ceraphant-T PTC31
meances, at different temperatures and feed concentrations, were (Endress + Hausser) (accuracy of about ±1 mbar). A schematic rep-
not obtained. The most detailed pervaporation study for the ethyl resentation of the pervaporation membrane unit is shown in Fig. 1.
lactate system was performed by Delgado and co-authors for binary The temperature was controlled by a thermostated bath (Lauda,
[35] and quaternary mixtures [36] using a commercial polymeric Germany) with ethylene glycol/water solution that flows through
membrane Pervap 2201 (from Sulzer Chemtech), but, only the the jackets of feed vessels 1 and 2; pressure was set at 2 bar apply-
separation problem was focused, and concentration and temper- ing an overpressure of helium to the system in order to prevent
ature polarization were not taken into account. Although most vaporization of feed mixture over the entire temperature range.
of studies involve the use of polymeric membranes that provide
good selectivity and flux, these membranes do not normally sup- 2.2.1. Pervaporation experiments
port reaction conditions (in terms of concentrations, temperature, The feed was charged into the feed vessel 1 (1 L capacity) and
and pH, among others) and, therefore, are not appropriated for heated to the desired temperature. In order to heat the whole sys-
applications under those conditions. Inorganic membranes, mainly tem (tubing and membranes) to that same temperature, the feed
those of silica and zeolites, present better stability under acidic mixture is after re-circulated over the membrane modules using
and high temperature conditions, and are the best alternative to a positive displacement diaphragm pump (Hydra Cell G-03, Wan-
perform the dehydration of a reaction medium. Since the commer- ner International), in the absence of vacuum on the permeate side.
cial microporous silica membrane (from Pervatech) revealed better When the steady state is reached, the pervaporation experiment
selectivity and water flux than the one from Pervap SMS (from starts by applying vacuum to the permeate side by means of a
Sulzer Chemtech) in the dehydration of aqueous mixtures con- vacuum pump (Boc Edwards, U.K.). During the whole run, all per-
taining acetone and isopropanol [4], Pervatech membrane will be meated vapour is condensed on two parallel glass cold traps filled
considered in this work. Moreover, for the dehydration of ethanol, with liquid nitrogen. Finally, the collected permeate was defrosted,
the Pervatech membrane proved to have high flux and selectivity weighted and analyzed. The duration of the experiment is con-
[37]. ditioned by the trade-off between ensuring the nearly constant
In this work, the synthesis of ethyl lactate in a continuous inte- feed composition and enough amount of permeate. To verify the
grated pervaporation membrane reactor, using the commercial assumption of constant feed composition samples were collected
tubular membrane with higher flux and high selectivity (Pervat- before and after each experiment.
ech) is, for the first time, implemented and evaluated. Furthermore,
in order to better characterize the pervaporation process and aim- 2.3. Analytical method
ing to describe the PVMR unit, the effect of feed pressure, flowrate,
temperature and composition on the pervaporation performance is All the samples were analysed in a gas chromatograph
studied using batch experiments (BP), testing the different binary (Chrompack 9100, Netherlands) using a fused silica capillary col-
mixtures involved in the synthesis of ethyl lactate (ethanol/water, umn (Chrompack CP-Wax 57 CB, 25 m × 0.53 mm ID, df = 2.0 ␮m) to
ethyl lactate/water and lactic acid/water). Additionally, a new separate the compounds, and a thermal conductivity detector (TCD
mathematical model accounting for the mass transport phenomena 903 A) to quantify them. The column temperature was programmed
under non-isothermal conditions is developed, and after applied for with a 1.5 min initial hold at 110 ◦ C, followed by a 50 ◦ C/min ramp
a better understanding and description of the ethyl lactate produc- up to 190 ◦ C and held for 8.5 min. The injector and detector tem-
tion by means of a PVMR. perature was maintained at 280 and 300 ◦ C, respectively. Helium
C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55 45

Fig. 1. Setup of the pervaporation membrane reactor unit.

N50 was used as the carrier gas with a flowrate of 10.5 mL/min. component i, and Pperm is the total pressure on the permeate side.
The species quantification using this analytical method presents The saturation pressure of pure components was estimated by the
an inherent average error less than 2%. Antoine equation [41].
Some works consider the temperature influence on the total
3. Pervaporation studies flux to measure an apparent activation energy [8,36,42]. However,
in this work, the activation energy of permeation is calculated by
The experimental results of the pervaporation studies for the an Arrhenius-type equation for the permeance temperature depen-
different binary mixtures (ethanol/water, ethyl lactate/water and dence [43]:
lactic acid/water) are presented in this section. The effect of the  
absolute feed pressure, feed flowrate, operation temperature, and −Eperm,i
Qmemb,i = Qmemb,0 exp (4)
water feed mole fraction onto the membrane performance is evalu- RT
ated. The membrane permeabilities were measured in the absence
of mass transfer limitations in the boundary layer like shown by where Qmemb,0 is the pre-exponential factor, Eperm,i is the activa-
the preliminary studies. tion energy of permeation, which is a combination of activation
energy of diffusion and the heat of adsorption on the membrane
3.1. Pervaporation transport (Eperm,i = ED,i + Hs ), T is the absolute temperature and R is the ideal
gas constant.
The pervaporation performance of the membrane was evaluated
in terms of pervaporation flux and separation factor (membrane
selectivity). The process separation factor (˛) is defined as: 3.2. Preliminary studies
yi xj
˛= (1) 3.2.1. Evaluation of the membrane quality
xi yj
The driving force for the transport through the membrane is
where xi is the liquid mole fraction of component i on the feed side based on the species chemical potential difference over the mem-
and yi is the mole fraction of component i on the permeate side. brane and, theoretically, the absolute feed pressure affects the
The partial flux of one component through the membrane is chemical potentials via the Poynting factor [44], which might affect
given by: either the flux or the selectivity. However, this influence is often
negligible since the Poynting factor is one at low pressures and,
Ji = wperm,i Jtot (2)
therefore, a way to detect membrane imperfections is by perform-
where Jtot is the total permeation flux expressed in kg/(m2 s) and ing pervaporation experiments at different feed pressures. If the
wperm,i is the mass fraction of component i on the permeate. permeate composition and the total flux remains constant over the
The solution-diffusion model [38] was successfully applied in studied pressure range the membrane quality can be, in principle,
the description of solvent dehydration using microporous silica guaranteed [45].
membranes [37,39]. This model provides the following transport The influence of the feed pressure onto the pervaporation mem-
equation for the permeation molar flux of a component through brane performance, is shown in Fig. 2, where the total permeation
the membrane: flux and permeate composition are given as a function of feed
pressure, keeping constant the remaining conditions (temperature,
Ji = Qmemb,i (xi i p0i − yi Pperm ) (3)
flowrate, water feed concentration, permeate pressure). The varia-
where Qmemb,i is the permeance of component i through the tions observed either on the permeate compositions or on the total
membrane (mol/(m2 s Pa)), which is equal to Pi /, being Pi the flux are within the experimental error, being possible to conclude
permeability coefficient of component i (mol/(m s Pa)), and  the that pervaporation is not affected by the feed pressure, confirming
thickness of the selective layer of the membrane,  i is the activity the membrane quality. Bruijn et al. [46] studying the effects of feed
coefficient (calculated by the UNIQUAC model using the parameters pressure on a similar methylated silica membrane also found the
determined in previous work [40]), p0i is the saturation pressure of same conclusion.
46 C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55

Fig. 2. Pressure dependence of total permeation flux and of permeates composition


for the system ethanol/water (T = 321.15 K, xW,F = 0.183 ± 0.001, Pperm = 20 mbar).
Fig. 3. Total permeation flux as a function of feed flowrate (T = 321.15 K,
xW,F = 0.182 ± 0.006, Pperm = 25 mbar).
3.2.2. Evaluation of mass transfer limitations in the boundary
layer
The effect of concentration polarization on the pervaporation 3.3. Detailed studies
process was studied for water/ethanol mixtures, at constant feed
water concentration and operating temperature, varying the feed The influence of feed composition, at different operating tem-
flowrate. Analyzing Fig. 3, it is possible to conclude that for feed peratures, on the total permeation flux and permeate composition
flowrates higher than 19 L/h, the total flux remains constant, indi- is shown in Figs. 4 and 5 for the water/ethanol and water/ethyl
cating the absence of mass transfer resistance from the bulk liquid lactate pairs, respectively.
phase to the feed membrane interface. The mass transfer effects in As can be seen, for both ethanol/water and ethyl lactate/water
the boundary layer must be evaluated at the highest working tem- systems, the total permeation flux increases with feed water com-
perature, since the flux through the membrane is higher. However, position and operation temperature (Fig. 4(a) and Fig. 5(a)). Like
it was observed that even at the highest temperature (343.15 K) already mentioned for dehydration in similar membranes [4,47]
the effect of concentration polarization is negligible at a flowrate of it is clear that the flux depends strongly on the temperature. For
19 L/h. Therefore, the remaining pervaporation experiments were example, at the same feed composition in the water/ethyl lactate
performed at a feed flowrate of 20 L/h. mixtures, it is observed an increase around 167% in the total per-

Fig. 4. Pervaporation performance for water/ethanol mixtures. (a) Influence of feed water mole fraction on total permeation flux at different operating temperatures. (b)
Influence of feed water mole fraction on permeate composition at different operating temperatures.

Fig. 5. Pervaporation performance for water/ethyl lactate mixtures. (a) Influence of feed water mole fraction on total permeation flux at different operating temperatures.
(b) Influence of feed water mole fraction on permeate composition at different operating temperatures.
C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55 47

Fig. 6. Influence of the temperature on the separation factor for the binary mixtures
water/ethanol and water/ethyl lactate.

Fig. 7. Temperature dependence of species permeances (mol/(s m2 Pa)).

meation flux when the operation temperature changes from 321 to


addition to the lowest vapor pressure, leads to infinite membrane
343 K. In fact, the permeation driving force increases with temper-
selectivity towards water in lactic acid mixtures.
ature for all species. Moreover, since water is the most permeated
component, an increase in the feed water content leads to a higher
3.4. Parameter estimation
water driving force, and, consequently, the total flux is significantly
enhanced. Regarding the permeate composition, it is almost invari-
The pervaporation process design requires the knowledge of the
able with temperature for both ethanol/water (Fig. 4(b)) and ethyl
permeance of each component as function of temperature (Eq. (4)).
lactate/water (Fig. 5(b)) systems; and with feed water content for
However, there is evidence that the feed water content affects per-
ethyl lactate/water pair (varies from 0.996 to 0.999). However, for
meance [36] and, therefore, both factors are needed to be accounted
water/ethanol pair, it is observed that the water permeate mole
for.
fraction increases with the feed water concentration.
For water/lactic acid mixtures two pervaporation experiments
3.4.1. Permeance temperature dependence
were performed with a feed water mole fraction of 0.77, at two
Experimental permeance for each species can be calculated
different temperatures (321.15 and 343.15 K). Observing that the
combining using the experimental information found so far in Eq.
permeate stream was pure water no more experiments were per-
(3), simultaneously with the activity coefficients calculated by the
formed for this binary mixture.
UNIQUAC model [40]. Those values are presented in Fig. 7.
The process separation factor for the pairs water/ethanol and
From the slope and intercept of the linear regression, the acti-
water/ethyl lactate is plotted as a function of the temperature in
vation energy of permeation and the pre-exponential factor are
Fig. 6. It can be seen that the membrane presents good selectivity
estimated for each species, like presented in Table 2. Although
towards water in both pairs, but the selectivity to water is higher
the permeation flux increases with temperature due to increasing
in mixtures with ethyl lactate most probably due to the higher
driving forces, the membrane permeability decreases as indicated
molecular size of ethyl lactate compared with that of ethanol, as
by the negative values of the activation energies of permeation
supported by the values of the gyration radius presented in Table 1
(Table 2). As already mentioned, the activation energy is the sum
which are an indication of the molecular sizes. It would then be
of the diffusion activation energy and the heat of adsorption on the
expected permeation of lactic acid through the membrane. How-
membrane, indicating that the permeation of water, ethanol and
ever, the molecular size is not the only factor affecting selectivity;
ethyl lactate is governed by the adsorption. Similar results were
the lower lactic acid vapor pressure (Table 1) significantly reduces
found for water/ethanol permeation through silica microporous
the permeation driving force for this component. Additionally, the
membranes [39].
affinity between species and membrane should be considered as
The error introduced considering composition invariant perme-
a secondary factor (the molecular diameter is the most impor-
ances can be calculated from the mean relative deviation (MRD),
tant). This affinity is directly related with the dipole moment of
defined by Eq. (5).
each component (Table 1): small dipole moment might complicate
⎛ ⎞
the penetration of molecules into the hydrophilic silica layer. Com-
1   Ji,calc − Ji,exp 
paring lactic acid with ethyl lactate properties, the molecular sizes MRD = ⎝   ⎠ × 100% (5)
are similar, but the first has the lowest dipole moment, which in nexp  Ji,exp 
nexp

In order to decrease the deviations between theoretical


Table 1 and experimental fluxes, parameter estimation considering also
Physical properties for the different species (water, ethanol, ethyl lactate and lactic
composition dependence will be addressed in the following sub-
acid).
section.
Molecule Radius of gyration Dipole moment Vapour pressure
(Å)a (Debyes)a (bar)b
Table 2
Water 0.615 1.8497 1.24 × 10−1 Pervaporation parameters of Eq. (4) and the mean relative deviation (MRD).
Ethanol 2.259 1.6908 2.94 × 10−1
Component Qmemb,0 (mol/(s m2 Pa)) Eperm (kJ/mol) MRD (%)
Ethyl lactate 3.622 2.4000 1.92 × 10−2
Lactic acid 3.298 1.1392 1.08 × 10−3 Ethanol 2.36 × 10−12 −22.60 14.51
a Ethyl lactate 1.99 × 10−9 −10.42 24.13
Data taken from Ref. [43].
b Water 2.97 × 10−8 −13.93 13.94
Vapour pressure values at 323.15 K.
48 C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55

Fig. 8. Water flux as a function of the driving force. Fig. 9. 1/kov,W versus 1/vbF , b = 0.97.

3.4.2. Permeance temperature and water content dependence co-authors [51]; and, less likely, swelling as already reported
According to Eq. (3), an average water permeance value is for MFI zeolites membranes in the presence of certain solvents
calculated from the slope of water flux as a function of water [52].
driving force. Fig. 8 shows a linear pattern for the experimen- The detailed expression of water permeance (Eq. (6)) leads to a
tal values obtained in this work, indicating a water permeance MRD between experimental and calculated values of 10.35%, which
of 4.72 × 10−6 mol/(m2 s Pa), which is very similar to the aver- is about 26% smaller than the one obtained when just the tem-
age of the calculated water permeances from the experimental perature influence was considered. Although derived from binary
data (4.76 × 10−6 ± 8.62 × 10−7 mol/(m2 s Pa)). This behaviour was mixture data, expression (6) can be applied to estimate the water
expected considering other works with ceramic silica membranes, permeance in quaternary mixtures, since it depends only on the
like the dehydration of isopropanol where the water permeance water content no matters the remaining constituents. This is in
was independent of the temperature and water feed concentration agreement with experimental pervaporation results by Delgado
[48]. However, despite the linear trend observed, there is certain et al. [36], who studied binary and quaternary mixtures involved
data dispersion, leading to deeper analysis of these results, where in the ethyl lactate synthesis. Therefore, it can be stated that the
the water permeance was calculated for each experiment by Eq. water transport is barely affected by the presence of the other com-
(3), decoupling the effects of temperature and composition (see ponents supporting the validity of Eq. (6). For ethanol and ethyl
Appendix). In consequence, the following empirical correlation was lactate, no relation was found between their permeances in the
obtained for water permeance as function of the temperature and membrane and the feed contents. Therefore, the parameters of
water feed content: Table 2 were used to determine the permeances of these species
at different operating temperatures and water feed contents.
Qmemb,W = 3.278 × 10−11 exp(18.64xW ) exp

50377x − 32326 3.4.3. Estimation of the boundary layer mass transfer coefficient
W
× − (mol/(s m2 Pa)) (6) (Kbl )
RT
In pervaporation processes, in addition to the permeation resis-
This expression, reflects the fact that water adsorption is the tance in the membrane there is the resistance in the boundary layer
controlling mechanism (negative values of activation energy of (concentration polarization). This can be conveniently represented
permeation), as sated before; but includes the influence of the by the resistance in series model, where the overall resistance to
water feed content. According to the solution-diffusion theory, the transport is the sum of boundary layer and membrane resistances.
activation energy for the permeation process is a combination of The resistance in series equation for pervaporation was derived by
activation energy of diffusion and heat of adsorption on the mem- Wijmans and collaborators [9]:
brane (Eperm,i = ED,i + Hs ). It is known that the first term is positive 1 1 i p0i Vmol,i
(ED,i ), and, for physisorption process, the second term (Hs ) is neg- = + (7)
kov,i Qmemb,i avbF
ative. Moreover, the heat of adsorption depends on the degree
of surface coverage, therefore, Hs is indeed an isosteric heat of in which kov,i is a global membrane mass transfer coefficient
adsorption. The membrane tested in this work has a modified (mol/(s m2 Pa)), that combines the resistance due to the diffusive
silica selective layer coated onto gamma alumina. Both materi- transport in the boundary layer with the membrane resistance,
als (silica and alumina) are hydrophilic and, therefore, commonly Vmol,i is the molar volume of component i (m3 /mol), Kbl = avbF where
used for water removal. It has been experimentally demonstrated vF is the feed liquid velocity (m/s). Eq. (7) will be used to determine
that the isosteric heat of water adsorption decreases significantly the mass transfer resistance in the boundary layer.
with the increase of the surface coverage for silica, alumina and The pervaporation experimental data obtained for water/
silica-alumina materials [49], and for silica gel [50]. This is in ethanol mixtures using different feed flowrates was used to plot the
agreement with the influence of water feed content on water inverse of the global membrane mass transfer coefficient (1/kov ) of
permeance described by Eq. (6): the isosteric heat of adsorption water and ethanol as a function of 1/vbF , as shown in Figs. 9 and 10,
decreases as the water feed content increases, leading to higher respectively. The value of parameter b was chosen in order to keep
activation energy of permeation (see Table A1), and consequently the data points on a straight line and also minimizing the difference
the water permeance decreases when feed water concentration between the inverse of the intercept of the line and the experimen-
increases. Other possible explanations for the behaviour exper- tal permeance measured in absence of mass transfer limitations on
imentally observed are a crowding (saturation) effect of water the boundary layer (being the best fit for b = 0.97). The parame-
diffusion through the silica pores, as proposed by Baker and ter a was obtained from the slope of this line. By this analysis it
C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55 49

i, through the membrane, defined as:


Ji = kov,i (xi i p0i − yi Pperm ) (13)
where kov,i is the global membrane mass transfer coefficient, that
combines the resistance due to the diffusive transport in the bound-
ary layer with the membrane resistance [9]:

1 1 i p0i Vmol,i
= + (14)
kov,i Qmemb,i Kbl

For laminar flow and Graetz number [dint 2 v/(D L)] much greater
m
than one, the mass transfer coefficient for transport in the boundary
layer, Kbl , is determined by the Lévêque correlation [53]:

d 0.33
int
Sh = 1.62Re0.33 Sc 0.33 (Re < 2300) (15)
L
where Sh = Kbl dint /Dm , Re = dint / and Sc = /(Dm ) are the Sher-
Fig. 10. 1/kov,Eth versus 1/vbF , b = 0.97.
wood, Reynolds and Schmidt numbers, respectively, Dm is the
solute diffusivity in the boundary layer, dint is the inside diameter
of the membrane, L is the membrane length,  is the density and
was possible to obtain the following expressions to calculate the
 is the viscosity. The accuracy of the Lévêque correlation to esti-
boundary layer mass transfer coefficients for both species (ethanol
mate the mass transfer coefficient in the laminar regime has been
and water):
validated experimentally by several works [11,54]. The prediction
Kbl,W (m/s) = 7.27 × 10−5 [vF (m/s)]0.97 (8) of the solute diffusivity was made using the Perkins and Geanko-
plis method [55]. Further details concerning its calculation can be
Kbl,Eth (m/s) = 3.56 × 10−7 [vF (m/s)]0.97 (9) found in a previous work [56].
Finally, the mole fraction of component i on the vapor phase
4. Modelling (permeate side), yi , is defined as:
J
4.1. Batch pervaporation model
yi = ni (16)
J
i=1 i

The mathematical model developed to describe the behaviour Fluid velocity variation in the membrane feed side calculated from the
of the batch pervaporation membrane (BPM) considers: total mass balance

dv  n

- Plug flow for the bulk fluid phase. = −Am Ji Vmol,i (17)
dz
- Total feed volume inside the tank and the retentate velocity vari- i=1
ations due to permeation of components. where n is the total number of components.
- Concentration polarization, where the resistance due to the dif-
fusive transport in the boundary layer is combined with the Retentate heat balance
membrane resistance in a global membrane resistance. 
n
∂T 
n
∂(T )
- Non-isothermal operation due to heat consumption for species Ĉp,i Cret,i + vĈp,i Cret,i + Am hF (T − Tm ) = 0 (18)
∂t ∂z
vaporization. i=1 i=1
- Temperature polarization.
where Ĉp,i is the liquid heat capacity of component i, T is the
absolute temperature in the feed side of the membrane, Tm is the
Following these assumptions the BPM model equations are:
membrane absolute temperature, and hF is the heat transfer coef-
Feed tank mass balance to component i ficient in the liquid boundary layer.

d(VCf,i ) Membrane heat balance


= Qret Cret,i − Qf Cf,i (10)
dt dint dint + ı 
n

2
hF (T − Tm ) = 2
HiV Ji (19)
where t is the time variable, V is the volume of the feed thank, Qf is 2
(rint + ı/2) − rint 2
(rint + ı/2) − rint
i=1
the flowrate fed to the membrane modules and Qret is the flowrate
at the end of the membrane modules, Cf and Cret are the liquid phase where rint is the internal radius of the membrane, ı is the membrane
concentration fed to the membrane and at the end of the membrane thickness, and HiV is the heat of vaporization of species i. The heat
modules, respectively. transport coefficient was estimated by the Sieder-Tate correlation,
valid for laminar pipe flow [57]:
Feed volume variation

dint
0.33
 0.14
dV b
= Qret − Qf (11) Nu = 1.86 Re Pr (20)
dt L w

Retentate mass balance to component i where Nu = hF dint / , Pr = Ĉp,i / are the Nusselt and Prandtl num-
bers, respectively, b and w are the viscosity of the liquid in the
∂Cret,i ∂(vCret,i ) feed and in the membrane wall, respectively, and is the thermal
+ + Am Ji = 0 (12)
∂t ∂z conductivity.
where z is the axial coordinate at the membrane modules, v is Initial and Danckwerts boundary conditions:
the superficial velocity, Am is the membrane area per unit mem-
brane modules volume and Ji is the permeate molar flux of species t=0: Cf,i = Cret,i = Ci,0 (21a)
50 C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55

V = V0 (21b) kinetic rate of the chemical reaction relative to the average particle
concentrations in the fluid phase given by [40]:
T = TF (21c)
aEth aLa − aEL aW /Keq
z=0: T = TF (22a) r = kc
2 (27)
NC
1+ K a
i=1 s,i i
v = vF (22b)

Cret,i = Cf,i (22c) where kc is the kinetic constant, Ks,i is the adsorption constant
for species i, Keq is the equilibrium reaction constant, and the
where subscripts 0 and F refer to initial state and membrane feed subscripts Eth, La, EL and W refer to ethanol, lactic acid, ethyl
conditions, respectively. lactate and water, respectively. Kinetic and thermodynamic param-
eters, needed in Eq. (27), were found in a previous work [40].
4.2. Pervaporation membrane reactor model The multi-component adsorption equilibrium isotherm is given
by:
A mathematical model was also developed to describe the
Qads,i Ki C̄p,i
behaviour of the tubular pervaporation membrane reactor that q̄i = n  (28)
similarly to the BPM model considers: 1+ K C̄
l=1 l p,i

where Qads,i and Ki represent the total molar capacity per unit vol-
- Concentration polarization, where the resistance due to the dif-
ume of the resin and the equilibrium constant for component i,
fusive transport in the boundary layer is combined with the
respectively. The adsorption parameters were also determined in a
membrane resistance in a global membrane resistance.
previous work [56].
- Temperature polarization.
Retentate heat balance
Additionally, it also takes in account:  ∂T 
n n
∂(T ) Am 
Ĉp,i Ci + uĈp,i Ci + hF (T − Tm ) + Hr b r = 0
- Axial dispersion flow for the bulk fluid phase, ∂t ∂z ε ε
i=1 i=1
- External and internal mass transfer for adsorbable species com- (29)
bined in a global particle resistance.
- Non-isothermal operation due to heat effects during species where Hr is the reaction enthalpy, which is 4.28 kJ/mol [40]. It
vaporization and reaction (slightly endothermic). must be mentioned that the heat of adsorption was neglected since
- Velocity variations due to permeation of components and adsorp- while some species are being adsorbed, others are being desorbed.
tion/desorption rates. The equation of the membrane heat balance is the one presented
- Constant column length and packing porosity. in the BPM model (Eq. (19)).
Initial and Danckwerts boundary conditions
Following this assumption the PVMR model equations are:
t=0: Ci = C̄p,i = Ci,0 , qi = qi,0 and T = TF (30)
Bulk fluid mass balance to component i: 
  ∂Ci 
z = 0 : uCi − Dax = uCi,F (31a)
∂Ci
+
∂(uCi )
+
(1 − ε) 3
K (C − C̄p,i ) = Dax

CT
∂xi

Am
J ∂z 
∂t ∂z ε rp L,i i ∂z ∂z ε i z=0

(23) u = u0 (31b)

where KL,i is the global mass transfer coefficient of the component T = TF (31c)
i, ε is the bed porosity, Dax , and u are the axial dispersion coefficient 
∂Ci 
and the interstitial velocity, respectively, rp is the particle radius, z = Lc : =0 (31d)
and C̄p is the average particle concentration. The calculation of the ∂z 
z=Lc
global mass transfer coefficient (KL ) is presented in detail in a previ- where F and 0 refer to the feed and initial states, respectively.
ous work [56], and the axial dispersion coefficient (Dax ) is estimated The viscosities, heat capacities, thermal conductivities and heats
from the empirical correlation [58] valid for liquids in packed beds: of vaporization of pure components were calculated for each
εPep = 0.2 + 0.011Rep0.48 (24) temperature according to the DIPPR 801 database [59]. The mix-
ture heat capacity was estimated assuming linear mole fraction
in which Pep = dp u/Dax and Rep = dp u/ are the Peclet and averages, while the mixture density and mixture thermal conduc-
Reynolds numbers relative to particle, respectively. tivity were estimated assuming linear mass fraction averages. The
The permeate flux of membrane is defined by Eq. (13) (BPM methods used to predict binary and multi-component mixture vis-
model). cosities were presented in detail in a previous work [56].
Interstitial fluid velocity variation calculated from the total mass bal-
ance 4.3. Numerical solution

(1 − ε) 3  Am 
n n
du The above model equations were solved numerically using the
=− kL,i Vmol,i (Ci − C̄p,i ) − Ji (25)
dz ε rp ε gPROMS-general PROcess Modelling System version: 3.0.3. The
i=1 i=1
mathematical model involves a system of partial and algebraic
Pellet mass balance to component i equations (PDAEs). The axial domain was discretized using third
order orthogonal collocation in finite elements method (OCFEM).
3 ∂C̄p,i ∂q̄ 
K (C − C̄p,i ) = εp + (1 − εp ) i − i b r(C̄p,i ) (26) Ten finite elements per module with two collocation points in each
rp L,i i ∂t ∂t 1−ε element were used. The system of ordinary differential and alge-
where i is the stoichiometric coefficient of component i, b is the braic equation (ODAEs) was integrated over time using the DASOLV
bulk density, εp the particle porosity, q̄i is the average adsorbed integrator implementation in gPROMS. For all simulations a toler-
phase concentration of species i in equilibrium with C̄p,i , and r is the ance equal to 10−5 was fixed.
C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55 51

Table 3
Parameters used in the simulations.

Parameters

Feed temperature 323.15 K


Feed composition xEth,F = 0.3546; xLA,F = 0.3424; xW,F = 0.3030
Feed flowrate 1.0 mL/min
Permeate pressure 10 mbar
Bed porosity 0.424
Bulk density 374.0 g/L
Length of the bed 100 cm
Internal diameter 0.7 cm

terms of temperature drop, experimentally it was not observed


any temperature variation for ethanol dehydration, while it has
dropped 2 ◦ C for the experiment with ethyl lactate. According to
the simulation it would be expected a decrease of 2 and 4 ◦ C for
ethanol and ethyl lactate dehydrations, respectively.

5.2. Pervaporation membrane reactor


Fig. 11. Experimental water permeation flux as a function of water permeation flux
determined by the BPM model with the Kbl estimated (Lévêque correlation) and Kbl The performance of the pervaporation membrane reactor unit,
determined from experimental data. packed with A15-wet resin in the lumen side of the tubular mem-
branes (where the active layer of the silica membranes is placed)
was evaluated by simulation. It was considered that the resin was
5. Results and discussion initially saturated with ethanol and then it was fed with a mixture of
ethanol and 85 wt.% lactic acid aqueous solution. The permeation
5.1. Batch pervaporation fluxes were estimated using the expressions previously obtained
in Section 3.4, and a compilation of the parameters used in the
Within the mathematical model proposed for the batch perva- simulations is presented in Table 3.
poration, the mass transfer coefficient in the boundary layer is, for The PVMR unit was evaluated considering isothermal and non-
all the species, estimated through the Lévêque correlation. How- isothermal operation and its performance was compared with the
ever, its accuracy should be validated experimentally. In Section one of a fixed bed reactor (FBR) in the same operational condi-
3.4.3, the boundary layer mass transfer coefficients, for water and tions. The concentration histories at the end of the PVMR and FBR,
ethanol, were determined from the experimental data (Eqs. (8) and considering isothermal operation, are shown in Fig. 13. It can be
(9)), and will be applied in the BPM model. The water permeation observed that the enhancement introduced by the water removal
fluxes were calculated either by using the experimental mass trans- through the membranes is significant for the ethyl lactate produc-
fer coefficient in the boundary layer (Eqs. (8) and (9)) or using the tion, where a 69% lactic acid conversion was achieved in the PVMR
Lévêque correlation (Eq. (15)). As can be observed in Fig. 11, there unit, at the steady state, while in the FBR the conversion obtained
is a good agreement between the water permeation flux estimated was just 29%.
using the experimental mass transfer coefficient with the one using Considering, alternatively, a non-isothermal operation (see
the Lévêque correlation, proving its accuracy in the prediction of Fig. 14(b)), the lactic acid conversion at steady state of the FBR is
the boundary layer mass transfer coefficients. Besides, small devi- about 29%, much better than 11% found for the PVMR (Fig. 14(a)).
ations are observed between the experimental fluxes and the ones This low performance of the PVMR is due to the temperature drop
determined by the BPM model. noticed in the bulk, around 36 ◦ C, while only 2 ◦ C was noticed for
In order to validate the BPM non-isothermal model, the dehy- the FBR (Fig. 14(a) and (b), respectively). This slightly decrease in
dration of ethanol and ethyl lactate aqueous solutions was studied. the FBR is due to the heat needed to the reaction only (endother-
The evolution of water mole fraction in the retentate stream is in mic reaction), while for the PVMR there is a large heat consumption
agreement with the experimental data: for both ethanol and ethyl to vaporize the species. Since the flowrate is very small, the heat
lactate cases the retentate composition after 15 and 10 min, respec- capacitance provided by the liquid stream is low and therefore
tively, is similar to the theoretical values, as shown in Fig. 12. In its temperature is drastically decreased, leading to lower reaction

Fig. 12. Evolution of water composition on retentate and temperature predicted by the non-isothermal BPM model: (a) dehydration of 81% of ethanol in aqueous solution
at 336 K; (b) dehydration of 44% of ethyl lactate in aqueous solution at 344 K.
52 C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55

Fig. 13. Concentration histories at the reactor outlet considering isothermal operation: (a) PVMR; (b) FBR.

Fig. 14. Concentration and temperature histories at the reactor outlet considering non-isothermal operation: (a) PVMR; (b) FBR.

kinetic rates, higher mass transfer resistances and lower water per- ties can be obtained, when this reactor is operated under isothermal
meation fluxes through the membranes. conditions.
Clearly, to operate the PVMR for efficient ethyl lactate produc- These results demonstrate that the PVMR has a great potential
tion, it is necessary to provide the heat required for vaporization for the ethyl lactate production, where high conversions and puri-
and reaction, operating in isothermal conditions. As stated before, ties can be obtained, when this reactor is operated under isothermal
the silica membranes used in this work have the selective layer conditions. Since PVMR integrates the reaction and separation into
coated inside the membrane tube. Therefore, the most suitable way a single device, it is necessary to find the operating conditions that
to supply heat to the retentate liquid is introducing a heated sweep satisfy, simultaneously, the equilibrium and kinetics of reaction and
gas in the permeate side instead of using vacuum. However, when the membrane permeation. Detailed models should be used for the
the membrane selective layer is coated on the external side of the appropriate design of a continuous tubular PVMR unit, from lab-
membrane tube (shell side), the heat can be supplied through an oratory to industrial scale, since simplified models might lead to
appropriate heated solution re-circulated through jacketed mod- under-sized units. As example, Fig. 16 shows the influence of the
ules; in this situation, the heat is rapidly transferred to the retentate space time on the lactic acid conversion, for two different ratios
liquid stream. between membrane length and diameter, with and without taking
Assuming that the PVMR unit could operate isothermally, it is into account the membrane concentration polarization effects. As
possible to determine by simulation the number of membranes can be seen, the concentration polarization effects decrease signif-
connected in series needed to maximize the lactic acid conversion
of the same feed processed before. For an operating temperature of
50 ◦ C (same conditions from Table 3), it is possible to achieve 93%
of lactic acid conversion and 84% of ethyl lactate purity at the PVMR
outlet at steady state (CEth = 0.65 mol/L; CLA = 0.62 mol/L; CEL =
7.67 mol/L; CW = 0.23 mol/L) when using a 250 cm long mem-
brane reactor (five tubular membranes, effective area of about
550 cm2 ). Trying to further enhance the separation performance,
as well as the reaction rate, the PVMR temperature was increased
to 70 ◦ C. In this case, it was obtained a lactic acid conversion of
98% and an ethyl lactate purity of 96% (CEth = 0.21 mol/L; CLA =
0.14 mol/L; CEL = 8.02 mol/L; CW = 0.02 mol/L). From the analy-
sis of the internal concentration profiles on the retentate side,
shown in Fig. 15, it is concluded that using only two membranes it
is possible to get about 90% of lactic acid conversion, but only 82%
of ethyl lactate purity; being necessary three more membranes to
maximize both lactic acid conversion and ethyl lactate purity.
These results demonstrate that the PVMR has a great potential
for the ethyl lactate production, where high conversions and puri- Fig. 15. Concentration profiles at steady state at 343.15 K.
C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55 53

non-isothermal operation due to heat consumption for species


vaporization; and (v) temperature polarization. The batch per-
vaporation membrane model was validated experimentally and,
therefore, it was extended to the integrated pervaporation mem-
brane reactor packed with the catalyst Amberlyst-15wet. This
model was used to evaluate the performance of the PVMR in both
isothermal and non-isothermal conditions. It was concluded that
non-isothermal operation worsens the performance of the PVMR,
being even worse than that obtained in the fixed bed reactor. In
isothermal conditions, the PVMR is a very attractive solution lead-
ing to a lactic acid conversion of 90% and ethyl lactate purity of 82%,
for the PVMR setup designed in this work (100 cm of length, two
membrane modules) operating at 70 ◦ C. For near lactic acid deple-
tion (98% conversion), it is produced ethyl lactate with 96% purity
if the membrane modules has 250 cm (five membranes in series),
as indicated by the model predictions. The PVMR model will be an
important tool in the design (membrane area, amount of catalyst),
Fig. 16. Lactic acid conversion at steady state as function of the space time with and and optimization, of an industrial pervaporation membrane reactor
without taking into account the concentration polarization effects, for a membrane
plant for the ethyl lactate production.
length of 1 m.

Acknowledgments
icantly the reactor performance. For L/D = 143 and a space time of
about 16 min, the concentration polarization effect reduces the lac-
The authors would like to thank Pervatech Company for the
tic acid conversion from 91 to 69%. Setting the conversion and feed
technical support. Carla S.M. Pereira gratefully acknowledges to
throughput, the simplified model leads to a membrane length half
Fundação para a Ciência e a Tecnologia (FCT, Portugal) the PhD
of the one really necessary.
Research Fellowships SFRH/BD/23724/2005. This work was also
supported by the project PPCDT/EQU/61580/2004 (FCT, Portugal).
6. Conclusions

Commercial hydrophilic silica membrane from Pervatech BV Appendix A. Supplementary data


(The Netherlands) was evaluated for the dehydration of ethanol,
ethyl lactate and lactic acid, in the temperature range from 48 to Supplementary data associated with this article can be found, in
72 ◦ C. First pervaporation studies indicate that the membranes have the online version, at doi:10.1016/j.memsci.2010.06.014.
no major imperfections since the total flux and water selectivity
is barely affect by absolute feed pressure. The influence of hydro-
dynamic conditions on the membrane polarization was analyzed, Nomenclature
and for velocity values higher than 0.14 m/s polarization effects are
eliminated. The effect of feed temperature and composition on the Am membrane area per unit membrane modules vol-
pervaporation performance was evaluated by batch experiments. ume (m2 /m3 )
It was concluded that the microporous silica membranes have high C liquid phase concentration (mol/m3 )
flux and high selectivity for water, while ethanol and ethyl lac- Cf liquid phase concentration in the thank (mol/m3 )
tate permeation is reduced and lactic acid does not permeate at C̄p average particle concentration (mol/m3particle )
all. In summary, the permeances for all species through the micro- Ĉp liquid heat capacity (J/(mol K))
porous silica membranes, as a function of temperature and feed Cret liquid phase concentration in the retentate (mem-
water content, are described by: brane feed side) (mol/m3 )
  Dax axial dispersion coefficient (m2 /s)
−12 22.60 × 103
Qmemb,Eth = 2.36 × 10 exp (mol/(s m2 Pa)) df film thickness (␮m)
RT
dint internal diameter of the membrane (m)
Dm solute diffusivity in the boundary layer (m2 /s)
Qmemb,LA = 0 (mol/(s m2 Pa)) ED activation energy of diffusion (J/mol)
  Eperm activation energy of permeation (J/mol)
10.42 × 103 hF heat transfer coefficient in the liquid boundary layer
Qmemb,EL = 1.99 × 10−9 exp (mol/(s m2 Pa)) (W/K)
RT
Hs heat of adsorption (J/mol)
HV heat of vaporization (J/mol)
Qmemb,W = 3.278 × 10−11 exp(18.64xW ) exp
Hr enthalpy of reaction (J/mol)

50377x − 32326 J permeate flux (mol/(m2 s))
W
× − (mol/(s m2 Pa))
RT Jtot total permeate flux (mol/(m2 s))
Kbl boundary layer mass transfer coefficient (m/s)
A mathematical model was developed for the batch perva- kc kinetic constant (mol/(kg s))
poration membrane, considering (i) plug flow for the bulk fluid Keq equilibrium reaction constant
phase; (ii) total feed volume inside the tank and retentate velocity Ki Langmuir equilibrium constant (m3 /mol)
inside the membrane variations due to permeation of compo- KL global mass transfer coefficient
nents; (iii) concentration polarization, where the resistance due kov global membrane mass transfer coefficient
to the diffusive transport in the boundary layer is combined with (mol/(m2 s Pa))
the membrane resistance in a global membrane resistance; (iv)
54 C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55

References

Ks adsorption constant in Eq. (27) [1] R.Y.M. Huang, R. Pal, G.Y. Moon, Crosslinked chitosan composite membrane
 thickness of the selective layer of the membrane (m) for the pervaporation dehydration of alcohol mixtures and enhancement of
structural stability of chitosan/polysulfone composite membranes, J. Membr.
L column length (m)
Sci. 160 (1999) 17–30.
n total number of components [2] V. Van Hoof, C. Dotremont, A. Buekenhoudt, Performance of Mitsui NaA type
Nu Nusselt zeolite membranes for the dehydration of organic solvents in comparison
P permeability coefficient (mol/(m s Pa)) with commercial polymeric pervaporation membranes, Sep. Purif. Technol. 48
(2006) 304–309.
p0 saturation pressure (s) [3] P.D. Chapman, T. Oliveira, A.G. Livingston, K. Li, Membranes for the dehydration
Pep Peclet number relative to particle of solvents by pervaporation, J. Membr. Sci. 318 (2008) 5–37.
Pperm total pressure on the permeate side (s) [4] C. Casado, A. Urtiaga, D. Gorri, I. Ortiz, Pervaporative dehydration of organic
mixtures using a commercial silica membrane: determination of kinetic param-
Pr Prandtl eters, Sep. Purif. Technol. 42 (2005) 39–45.
q̄ solid phase concentration in equilibrium with the [5] L. Lin, Y. Zhang, Y. Kong, Recent advances in sulfur removal from gasoline by
fluid concentration inside the particle (mol/m3 ) pervaporation, Fuel 88 (2009) 1799–1809.
[6] B. Smitha, D. Suhanya, S. Sridhar, M. Ramakrishna, Separation of
Qads total molar capacity per unit volume of resin organic–organic mixtures by pervaporation – a review, J. Membr. Sci.
(mol/m3 ) 241 (2004) 1–21.
Qmemb permeance (mol/(m2 s Pa)) [7] T. Ohshima, T. Miyata, T. Uragami, Selective removal of dilute benzene from
water by various cross-linked poly(dimethylsiloxane) membranes containing
Qmemb,0 pre-exponential factor (mol/(m2 s Pa)) tert-butylcalix[4]arene, Macromol. Chem. Phys. 206 (2005) 2521–2529.
Qf flowrate fed to the membrane modules (m3 /s) [8] M. Khayet, C. Cojocaru, G. Zakrzewska-Trznadel, Studies on pervaporation sep-
Qret flowrate at the exit of the membrane modules aration of acetone, acetonitrile and ethanol from aqueous solutions, Sep. Purif.
Technol. 63 (2008) 303–310.
(m3 /s)
[9] J.G. Wijmans, A.L. Athayde, R. Daniels, J.H. Ly, H.D. Kamaruddin, I. Pinnau, The
r rate of reaction (mol/(kg s)) role of boundary layers in the removal of volatile organic compounds from
R ideal gas constant (J/(mol K)) water by pervaporation, J. Membr. Sci. 109 (1996) 135–146.
rint internal radius of the membrane (m) [10] I. Ortiz, A. Urtiaga, R. Ibáñez, P. Gómez, D. Gorri, Laboratory- and pilot plant-
scale study on the dehydration of cyclohexane by pervaporation, J. Chem.
rp particle radius (m) Technol. Biotechnol. 81 (2006) 48–57.
dp particle diameter (m) [11] R.O. Crowder, E.L. Cussler, Mass transfer resistances in hollow fiber pervapora-
Re Reynolds tion, J. Membr. Sci. 145 (1998) 173–184.
[12] C. Lipski, P. Côté, The use of pervaporation for the removal of organic contam-
Rep Reynolds number relative to particle inants from water, Environ. Prog. 9 (1990) 254–261.
Sc Schmidt [13] C. Dotremont, S. Van den Ende, H. Vandommele, C. Vandecasteele, Concen-
Sh Sherwood tration polarization and other boundary layer effects in the pervaporation of
chlorinated hydrocarbons, Desalination 95 (1994) 91–113.
t time variable (s) [14] S. Bandini, A. Saavedra, G.C. Sarti, Vacuum membrane distillation: experiments
T absolute temperature in the feed side of the mem- and modeling, AIChE J. 43 (1997) 398–408.
brane (K) [15] A.M. Urtiaga, E.D. Gorri, J.K. Beasley, I. Ortiz, Mass transfer analysis of the per-
vaporative separation of chloroform from aqueous solutions in hollow fiber
Tm absolute membrane temperature (K) devices, J. Membr. Sci. 156 (1999) 275–291.
u interstitial velocity (m/s) [16] V. Gekas, B. Hallstrom, Mass transfer in the membrane concentration polariza-
V volume of the feed thank (m3 ) tion layer under turbulent cross flow. I. Critical literature review and adaptation
of existing sherwood correlations to membrane operations, J. Membr. Sci. 30
v superficial velocity (m/s)
(1987) 153–170.
Vmol molar volume (m3 /mol) [17] F. Lipnizki, R.W. Field, Mass transfer performance for hollow fibre modules
w mass fraction with shell-side axial feed flow: using an engineering approach to develop a
xi liquid mole fraction of component i in the feed side framework, J. Membr. Sci. 193 (2001) 195–208.
[18] T.A.C. Oliveira, U. Cocchini, J.T. Scarpello, A.G. Livingston, Pervaporation mass
yi mole fraction in the vapor phase of component i transfer with liquid flow in the transition regime, J. Membr. Sci. 183 (2001)
z axial coordinate at the membrane modules (m) 119–133.
[19] E. Favre, Temperature polarization in pervaporation, Desalination 154 (2003)
129–138.
Greek letters [20] H.O.E. Karlsson, G. Trägardh, Heat transfer in pervaporation, J. Membr. Sci. 119
i activity coefficient of component i (1996) 295–306.
ε bulk porosity [21] Y. Zhu, R.G. Minet, T.T. Tsotsis, A continuous pervaporation membrane reactor
for the study of esterification reactions using a composite polymeric/ceramic
εp particle porosity membrane, Chem. Eng. Sci. 51 (1996) 4103–4113.
i stoichiometric coefficient of component i [22] T.A. Peters, N.E. Benes, J.T.F. Keurentjes, Zeolite-coated ceramic pervaporation
b bulk density (kg/m3 bed) membranes; pervaporation–esterification coupling and reactor evaluation,
Ind. Eng. Chem. Res. 44 (2005) 9490–9496.
 fluid density (kg/m3 )
[23] B.G. Park, T.T. Tsotsis, Models and experiments with pervaporation membrane
viscosity (Pa s) reactors integrated with an adsorbent system, Chem. Eng. Process. 43 (2004)
ı membrane thickness (m) 1171–1180.
thermal conductivity (W/(m K)) [24] S.Y. Lim, B. Park, F. Hung, M. Sahimi, T.T. Tsotsis, Design issues of perva-
poration membrane reactors for esterification, Chem. Eng. Sci. 57 (2002)
˛ process separation factor 4933–4946.
[25] X. Feng, R.Y.M. Huang, Studies of a membrane reactor: esterification facilitated
Subscripts by pervaporation, Chem. Eng. Sci. 51 (1996) 4673–4679.
[26] M.T. Sanz, J. Gmehling, Esterification of acetic acid with isopropanol coupled
F feed with pervaporation. Part II. Study of a pervaporation reactor, Chem. Eng. J. 123
i component i (i = Eth, La, EL, W) (2006) 9–14.
0 initial conditions [27] Ó. de la Iglesia, R. Mallada, M. Menéndez, J. Coronas, Continuous zeolite mem-
brane reactor for esterification of ethanol and acetic acid, Chem. Eng. J. 131
Perm permeate (2007) 35–39.
Eth ethanol [28] P. Gómez, R. Aldaco, R. Ibáñez, I. Ortiz, Modeling of pervaporation pro-
La lactic acid cesses controlled by concentration polarization, Comput. Chem. Eng. 31 (2007)
1326–1335.
EL ethyl lactate
[29] D.J. Benedict, S.J. Parulekar, S.P. Tsai, Esterification of lactic acid and ethanol
W water with/without pervaporation, Ind. Eng. Chem. Res. 42 (2003) 2282–2291.
[30] D.J. Benedict, S.J. Parulekar, S.P. Tsai, Pervaporation-assisted esterification of
lactic and succinic acids with downstream ester recovery, J. Membr. Sci. 281
(2006) 435–445.
C.S.M. Pereira et al. / Journal of Membrane Science 361 (2010) 43–55 55

[31] K. Wasewar, S. Patidar, V.K. Agarwal, Esterification of lactic acid with ethanol [45] M. Pera-Titus, J. Llorens, F. Cunill, On a rapid method to characterize intercrys-
in a pervaporation reactor: modeling and performance study, Desalination 243 talline defects in zeolite membranes using pervaporation data, Chem. Eng. Sci.
(2009) 305–313. 63 (2008) 2367–2377.
[32] D. Rathin, T. Shih-Perng, Esterification of fermentation-derived acids via per- [46] F. Bruijn, J. Gross, Z. Olujic, P. Jansens, F. Kapteijn, On the driving force of
vaporation, US Patent No. 57236391998. methanol pervaporation through a microporous methylated silica membrane,
[33] J.J. Jafar, P.M. Budd, R. Hughes, Enhancement of esterification reaction yield Ind. Eng. Chem. Res. 46 (2007) 4091–4099.
using zeolite A vapour permeation membrane, J. Membr. Sci. 199 (2002) [47] J.E. ten Elshof, C.R. Abadal, J. Sekulic, S.R. Chowdhury, D.H.A. Blank, Transport
117–123. mechanisms of water and organic solvents through microporous silica in the
[34] K. Tanaka, R. Yoshikawa, C. Ying, H. Kita, K.I. Okamoto, Application of zeolite T pervaporation of binary liquids, Micropor. Mesopor. Mater. 65 (2003) 197–208.
membrane to vapor-permeation-aided esterification of lactic acid with ethanol, [48] A.W. Verkerk, P. Van Male, M.A.G. Vorstman, J.T.F. Keurentjes, Description of
Chem. Eng. Sci. 57 (2002) 1577–1584. dehydration performance of amorphous silica pervaporation membranes, J.
[35] P. Delgado, M.T. Sanz, S. Beltrán, Pervaporation study for different binary mix- Membr. Sci. 193 (2001) 227–238.
tures in the esterification system of lactic acid with ethanol, Sep. Purif. Technol. [49] T. Morimoto, M. Nagao, J. Imai, Interaction between water and hydroxylated
64 (2008) 78–87. surface of silica-alumina, Bull. Chem. Soc. Jpn. 47 (1974) 2994–2997.
[36] P. Delgado, M.T. Sanz, S. Beltrán, Pervaporation of the quaternary mixture [50] A. Chakraborty, B.B. Saha, S. Koyama, Adsorption thermodynamics of silica gel-
present during the esterification of lactic acid with ethanol, J. Membr. Sci. 332 water systems, J. Chem. Eng. Data 54 (2009) 448–452.
(2009) 113–120. [51] R.W. Baker, J.G. Wijmans, Y. Huang, Permeability, permeance and selectivity: a
[37] S. Sommer, T. Melin, Performance evaluation of microporous inorganic mem- preferred way of reporting pervaporation performance data, J. Membr. Sci. 348
branes in the dehydration of industrial solvents, Chem. Eng. Process. 44 (2005) (2010) 346–352.
1138–1156. [52] M. Yu, J.L. Falconer, R.D. Noble, Characterizing non-zeolitic pore volume in zeo-
[38] J.G. Wijmans, R.W. Baker, The solution-diffusion model: a review, J. Membr. Sci. lite membranes by temperature-programmed desorption, Micropor. Mesopor.
107 (1995) 1–21. Mater. 113 (2008) 224–230.
[39] Y. Ma, J. Wang, T. Tsuru, Pervaporation of water/ethanol mixtures [53] M.A. Lévêque, Les lois de transmission de chaleur par convection, Ann. Mines
through microporous silica membranes, Sep. Purif. Technol. 66 (2009) 13 (1928) 201.
479–485. [54] S.R. Wickramasinghe, M.J. Semmens, E.L. Cussler, Mass transfer in various hol-
[40] C.S.M. Pereira, S.P. Pinho, V.M.T.M. Silva, A.E. Rodrigues, Thermodynamic equi- low fiber geometries, J. Membr. Sci. 69 (1992) 235–250.
librium and reaction kinetics for the esterification of lactic acid with ethanol [55] L.R. Perkins, C.J. Geankoplis, Molecular diffusion in a ternary liquid system with
catalyzed by acid ion exchange resin, Ind. Eng. Chem. Res. 47 (2008) 1453–1463. the diffusing component dilute, Chem. Eng. Sci. 24 (1969) 1035–1042.
[41] C.L. Yaws, Chemical Properties Handbook, McGraw-Hill, New York, 1999. [56] C.S.M. Pereira, V.M.T.M. Silva, A.E. Rodrigues, Fixed bed adsorptive reactor for
[42] C.S. Slater, T. Schurmann, J. MacMillian, A. Zimarowski, Separation of diacteone ethyl lactate synthesis: experiments, modelling and simulation, Sep. Sci. Tech-
alcohol–water mixtures by membrane pervaporation, Sep. Sci. Technol. 41 nol. 44 (2009) 2721–2749.
(2006) 2733–2753. [57] J. Welty, C.E. Wicks, G.L. Rorrer, R.E. Wilson, Fundamentals of Momentum, Heat
[43] X. Feng, R.Y.M. Huang, Liquid separation by membrane pervaporation: a review, and Mass Transfer, John Wiley & Sons, New York, 2008.
Ind. Eng. Chem. Res. 36 (1997) 1048–1066. [58] J.B. Butt, Reaction Kinetics and Reactor Design, Prentice-Hall, Englewood Cliffs,
[44] J.M. Smith, H.C.V. Ness, Introduction to Chemical Engineering Thermodynam- NJ, 1980.
ics, 4th ed., McGraw-Hill Book Co., New York, 1987. [59] DIPPR, Thermophysical Properties Database, 1998.

Potrebbero piacerti anche