You are on page 1of 18

Review

Received: 12 September 2019 Revised: 29 October 2019 Accepted article published: 6 November 2019 Published online in Wiley Online Library: 30 November 2019

(wileyonlinelibrary.com) DOI 10.1002/jctb.6264

Review of pervaporation and vapor


permeation process factors affecting the
removal of water from industrial solvents
Leland M Vane*

ABSTRACT
A recent review article (Journal of Chemical Technology & Biotechnology 94: 343–365 (2019)) identified several commercially-
available permselective materials for drying organic solvents with pervaporation (PV) and vapor permeation (V·P) separation
processes. The membrane materials included polymeric and inorganic substances exhibiting a range in the performance char-
acteristics: water permeance, water/solvent selectivity, and maximum use temperature. This article provides an overview of the
factors affecting the design of PV/V·P processes utilizing these membranes to remove water from common organic solvents.
Properties of the specific membrane and of the solvent substantially affect the PV/V·P separation. Equally important is the
impact of operating parameters on the overall separation. To study these impacts, simplified process performance equations
and detailed spreadsheet calculations were developed for single-pass and recirculating batch PV systems and for single-pass V·P
systems. Estimates of membrane area, permeate concentration, solvent recovery, permeate condenser temperatures, and heat-
ing requirements were calculated. Process variables included: solvent type, water permeance, water/solvent selectivity, initial
and final water concentrations, operating temperature (PV) or feed pressure (V·P), temperature drop due to evaporation (PV) or
feed-side pressure drop (V·P), and permeate pressure. The target solvents considered were: acetonitrile, 1-butanol, N,N-dimethyl
formamide, ethanol, methanol, methyl isobutyl ketone, methyl tert-butyl ether, tetrahydrofuran, acetone, and 2-propanol.
Published 2019. This article is a U.S. Government work and is in the public domain in the USA.

Supporting information may be found in the online version of this article.

Keywords: solvent reclamation; membrane processes; dehydration; pervaporation; vapor permeation; membrane-based separation

INTRODUCTION sodium Linde Type A (NaA) zeolites, chabazite (CHA) zeolites,


The life cycle environmental impact of using organic solvents can T-type zeolites, or hybrid silicas being commercially available. The
be reduced by recovering and reusing the solvents, avoiding the objective of this work is to provide an overview of the factors that
impacts associated with producing the virgin solvents and dispos- affect the design of PV/V·P processes that utilize these membranes
ing of the spent solvents.1–7 Solvent reuse/reprocessing requires to remove water from the target solvents.
separation technologies to recover and purify the solvents from Briefly, PV and V·P, depicted in Fig. 1, involve the sorption and
mixtures with other materials they encounter in industrial appli- diffusion of molecules in and through a thin, dense, permselec-
cations. For hydrophilic solvents, and even some solvents not tive membrane from a feed fluid into a permeate vapor. The dif-
typically considered ‘water loving’, water is often a contaminant ferences in sorption and diffusion of the molecules present in the
that must be removed prior to reuse. Unfortunately, many sol- mixture result in a separation: the composition of the stream that
vents form azeotropic mixtures with water, a property that makes is transported through the membrane (the ‘permeate’) is different
separation by evaporation or simple distillation complicated. than the composition of the fluid on the feed side of the mem-
Alternative separation technologies, including the membrane pro- brane. The composition of the stream rejected by the membrane
cesses of pervaporation (PV) and vapor permeation (V·P) have the (the ‘retentate’) is also altered. The driving force for transport is
potential to remove water from solvents, even under azeotropic the chemical potential difference between the feed-side mixture
conditions.8 A recent article by this author provides a state-of- and the permeate, represented by the fugacity or, more conve-
the-science review of materials available as the selective layer(s) niently, the partial pressure of each compound. In the context of
of PV and V·P membranes for the removal of water from ten com-
monly used organic solvents that are promising targets for recov-
ery and reuse.9 Those solvents included acetonitrile, 1-butanol, ∗ Correspondence to: LM Vane, US Environmental Protection Agency, 26 W.
N,N-dimethyl formamide (DMF), ethanol, methanol, methyl Martin Luther King Dr., Cincinnati, OH 45268 USA. E-mail: vane.leland@epa.gov
isobutyl ketone (MIBK), methyl tertiary butyl ether (MtBE), tetrahy-
Correction added on 21 January 2020, after first online publication: the
drofuran (THF), acetone, and 2-propanol (isopropanol). The analy-
supporting information files has been updated.
sis concluded that a robust and varied PV/V·P membrane industry
has developed, with membranes containing water-selective layers Center for Environmental Solutions and Emergency Response, US Environmen-
495

of poly(vinyl alcohol), polyimides, amorphous perfluoro polymers, tal Protection Agency, Cincinnati, OH, USA

J Chem Technol Biotechnol 2020; 95: 495–512 www.soci.org Published 2019. This article is a U.S. Government work
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org LM Vane

Figure 1. Illustrations of (a) vapor permeation (V·P) and (b) pervaporation (PV) processes with molar feed flow rate (N), ̇ species i feed fluid mole fraction
(x i ), feed temperature (T F ), feed pressure (pF ), permeate pressure (pP ), permeate mole fraction (yi ), and membrane permeance (Pi /𝓁). The mass transport
model for pervaporation assumes a theoretical vapor phase is present in equilibrium with the feed-side liquid yielding hypothetical feed vapor mole
fraction (xi∗ ) and total feed vapor pressure (pF* ).

Figure 2. Simplified schematic diagram of a pervaporation (PV) process with the option of returning the water-depleted retentate liquid to the feed
source tank (recirculating batch mode) or managing the retentate as the solvent product (single-pass mode). Common ancillary components, such as a
feed-retentate recuperative heat exchanger, are not shown.

this work on solvent dehydration, water preferentially permeates the V·P system and then a condenser to recover the retentate as a
the membranes relative to the organic solvents, leaving a reten- liquid. The vapor generation unit could be a simple complete evap-
tate depleted in water and a permeate enriched in water relative orator (as shown in Fig. 3) or a distillation unit – taking advantage
to the feed mixture. Thus, the retentate is the solvent to be reused of the stripping and enrichment capabilities of distillation units
while the permeate is either disposed of or processed for addi- as well as the ability of distillation units to handle dissolved and
tional resource recovery. suspended solids.10 Example flow diagrams of PV and V·P systems
Simplified example flow diagrams for PV and V·P are illustrated in combined with a batch evaporator, evaporator with a rectification
Figs 2 and 3, respectively. The distinction between PV and V·P is the column, and full distillation columns are illustrated in Figs S2–S7
phase of the feed-side fluid: liquid for PV and vapor for V·P. In PV, of the Supporting Information.
the water/solvent liquid mixture is heated to a target feed temper- Evaporation-based separation technologies, especially distilla-
ature, processed through a membrane module or set of modules tion, predominate in the arena of solvent recovery due to the usu-
as a liquid, and then the reduced-water solvent retentate liquid is ally advantageous vapor–liquid equilibrium (VLE) behavior, rela-
either returned to the feed source for a ‘batch’ operation (Option 1 tive ease of design, and potential for reduced incremental cost at
in Fig. 2) or considered the solvent product for a ‘single-pass’ oper- larger scale. The ratio of vapor phase concentrations divided by the
ation (Option 2 in Fig. 2). The permeate is a vapor, usually involving ratio of liquid phase concentrations when the two phases are in
vacuum conditions where the vacuum is generated through the equilibrium is referred to as ‘relative volatility’. Relative volatility of
combination of a vapor condenser and a small vacuum pump to species 1 compared to species 2 in a mixture, 𝛼12 VLE
, is a measure of
remove non-condensing gases, the latter originating from vacuum the separation produced by a single VLE stage:
leaks or dissolved gases. The flow diagram for V·P looks very similar
to that of PV. However, since most solvent-use applications require
the solvent in liquid form, employing V·P for solvent drying would (𝜔V1 ∕𝜔V2 ) (Y1 ∕Y2 ) 𝛾1 psat
1
𝛼12
VLE
= = = (1)
496

necessitate an evaporation step to create the vapor phase feed to (𝜔L1 ∕𝜔L2 ) (X1 ∕X2 ) 𝛾2 psat
2

wileyonlinelibrary.com/jctb Published 2019. This article is a U.S. Government work J Chem Technol Biotechnol 2020; 95: 495–512
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Review of pervaporation and vapor permeation process factors www.soci.org

Figure 3. Simplified schematic diagram of a vapor permeation (V·P) process with the option of returning the water-depleted retentate condensate to the
feed source tank (recirculating batch mode) or managing the retentate as the solvent product (single-pass mode).

where 𝜔ki is the mass fraction in phase k [either the liquid (L) or the
vapor (V)], X i and Y i are the mole fractions in the liquid and vapor,
respectively, 𝛾 i is the activity coefficient, and psati
is the saturated
vapor pressure, all of species i. The right side of the equation results
by inserting the VLE equilibrium relationship: Yi = Xi 𝛾i psat i
.
When species 1 is enriched in the vapor relative to species 2,
𝛼12
VLE
is greater than 1. Values of 𝛼12 VLE
less than 1 indicate the
opposite. When multiple ideal VLE stages are strung together, as in
a distillation column, the vapor composition ratio at the Nth stage
is calculated from the liquid composition ratio at the first stage and
the product of the relative volatility values for the N stages as:
( ) ( ) [ ]
Y1 X1 ∏
N
VLE
= (𝛼12 )n (2)
Y2 N X2 1 n=1

While the minimum reflux ratio required for a given separation VLE ) as function
Figure 4. Relative volatility of water/solvent mixtures (𝛼ws
can be calculated based on the Underwood method,11 the mini- of water weight fraction in the liquid phase (wwL ). Values calculated with

mum number of stages required to yield a specific top vapor prod- NRTL model at the normal boiling point temperature of each solvent.
VLE = 1) added as a reference. Values for
Solid gray ‘no separation’ line (𝛼ws
uct and bottom liquid product from a distillation column can be
1-butanol, methyl isobutyl ketone (MIBK), and methyl tertiary butyl ether
estimated with the Fenske equation:12 (MtBE) shown in the range of solubility of each with water.53–55
[( ) ( )]
ln Y1∕ X
Y2 N 2∕X1 1
Nmin = VLE
(3) distillation line,13 the minimum energy demand can be calculated
ln 𝛼 12 with the aid of pinch theory.14,15
The calculated volatility of water relative to that of a solvent in
VLE
where 𝛼 12 is the geometric mean of 𝛼12 VLE
over all stages, sometimes a binary mixture with each of the ten organic solvents highlighted
approximated as the geometric mean of 𝛼12 VLE
at the top and in this work is shown in Fig. 4 as a function of the weight fraction of
bottom of the column. Thus, larger relative volatility values result water in the liquid phase over the full concentration range. MIBK,
in fewer stages of VLE required to accomplish a desired separation. MtBE, and 1-butanol are only partially miscible with water and the
Conversely, as 𝛼12 VLE
approaches 1, the number of stages becomes curves in Fig. 4 are only shown in the regions where the liquid is a
infinite. If 𝛼12
VLE
drops below 1, the separation reverses with species single phase. The point at which a relative volatility curve crosses
2 enriched in the vapor instead of species 1. When 𝛼12 VLE
is equal to VLE
the ‘no separation’ line (𝛼ws = 1) is the azeotrope of that mixture.
1, there is no difference between the vapor phase and the liquid Several observations can be made from Fig. 4. First, the relative
phase concentrations, thus evaporation offers no separation of volatility curves for water in acetone, DMF, and methanol do
the two species. This condition is referred to as an azeotrope. To not cross the no separation line, consistent with the fact that
separate azeotropic mixtures, the VLE behavior must be altered these solvents do not form an azeotrope with water. In particular,
by changing conditions, such as adding a third compound, or an the water/acetone and water/methanol systems exhibit sol-
alternative technology must be introduced. While the minimum vent enrichment in the vapor phase (𝛼ws VLE
< 1) over the entire
497

number of stages can be estimated by determining the course of a concentration range. Although, at low water concentrations,

J Chem Technol Biotechnol 2020; 95: 495–512 Published 2019. This article is a U.S. Government work wileyonlinelibrary.com/jctb
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org LM Vane

𝛼ws
VLE
for water/acetone is practically equal to 1 and 𝛼ws VLE
for parameter. Common units for permeance are gas permeation units
water/methanol is only 0.45. Further, only the water/DMF system (GPU), defined as 10−6 cm3 (STP) cm−2 s−1 cmHg−1 with 1 kmol
exhibits water enrichment (𝛼ws VLE
> 1) in the vapor over the entire m−2 s−1 kPa−1 = 2.99 × 109 GPU. Common units for permeability
concentration range. Second, except for DMF, the solvents are are Barrer, defined as 10−10 cm3 (STP) cm cm−2 s−1 cmHg−1 with 1
strongly enriched in the vapor phase at low solvent concentra- kmol m m−2 s−1 kPa−1 = 2.99 × 1015 Barrer.
VLE The separation factor for PV or V·P is defined similarly to relative
tions with solvent/water relative volatility (i.e. 1∕𝛼ws ) ranging from
9.5 to 570 at infinite solvent dilution (i.e. the right edge of Fig. 4). volatility, as a ratio of the permeate vapor compositions to a ratio
Thus, evaporation processes effectively strip solvents from binary of the feed-side compositions for two species:
aqueous mixtures into the vapor phase at low solvent concentra-
tions. Third, for most of the solvents, the evaporative separation (w1P ∕w2P ) (y1 ∕y2 ) (j ∕j )
𝛽12 = = = 1 2 (6)
of mixtures containing less than 20 wt% water is challenging due (w1F ∕w2F ) (x1 ∕x2 ) (x1 ∕x2 )
to the presence of an azeotrope or low relative volatility – or
both. The relative volatility behavior in this water concentration where wik is the mass fraction of species i in stream k [feed (F),
range is highlighted in Fig. S1 in the Supporting Information. The retentate (R), or permeate (P)], respectively, and x i and yi are
water/ethanol system presents particularly significant distillation the mole fractions in the feed fluid and permeate vapor, respec-
challenges in this range because the relative volatility is close to 1 tively. Lowercase x and y are used for PV or V·P to distinguish
and exhibits an azeotrope at 4 wt% water. them from the VLE uppercase variables X and Y. The ratio of per-
Due to these observations, this article will focus on the removal meate mass fractions is equivalent to the ratio of mass fluxes
of water from mixtures containing less than 20 wt% water, using and, as shown in Eqn (6), the ratio of permeate mole fractions
the water/ethanol system as the principal example. Although is equal to the ratio of molar fluxes. Thus, given the feed com-
binary systems with a low relative volatility or an azeotrope cannot position, the separation factor can be calculated from the per-
easily be separated by simple distillation, the same is not true meate water purity and vice versa. Similarly, given the permeate
for separation processes that operate on principles other than composition, one can calculate total flux from water flux and vice
VLE. For example, PV and V·P leverage the sorption and diffusion versa.
differences of the mixture components in a dense membrane For both PV and V·P, the permeate is a reduced pressure vapor
material to separate the mixture. Because of this, even azeotropic where the partial pressure of a compound in the permeate, pPi ,
mixtures can be separated with PV or V·P with an appropriate is simply calculated as the permeate mole fraction (yi ) times the
membrane material and operating conditions. total permeate pressure (i.e. pPi = yi pP ). This is also the case for V·P
feed-side streams with partial pressures calculated from the mole
PV and V·P membrane performance descriptors fraction x i and total feed-side pressure pF (i.e. pFi = xi pF ). However,
In order to probe the variables that affect PV or V·P process per- for PV, the feed-side partial pressure of a compound experienced
formance, the relationships and parameters that define mem- by the membrane is calculated as if there was a hypothetical vapor
brane performance should be established. The performance of a phase (designated with an asterisk*) in equilibrium with both the
PV/V·P membrane for a given feed composition and temperature feed-side liquid and the membrane.16 The hypothetical feed-side
is most commonly reported in terms of total mass flux, or mass partial vapor pressure (pF∗i
) is calculated from the actual feed-side
flux of water, and either separation factor of species 1 relative to liquid composition (x i ) using VLE behavior as:
species 2 (𝛽 12 ) or permeate purity in terms of water weight frac-
tion (wwP ). The mass flux of a species (Ji , in kg m−2 s−1 ) is defined pF∗
i
= xi 𝛾i psat
i
(7)
as the mass of a species (mi , in kilograms) permeating the mem-
brane per unit membrane area (A, in m2 ) and unit time (t, in Such a calculation requires either experimental VLE data or
seconds) as: thermodynamic models to calculate the activity coefficient (𝛾 i )
m for each compound in a solvent/water mixture and the saturated
Ji = i (4) vapor pressure (psat ) of each compound. The activity coefficients
At i
calculated with the Non-Random Two-Liquid (NRTL) model for
The molar flux, ji (in kmol m−2 s−1 ), is simply the mass flux divided water and solvents in binary mixtures at the normal boiling point
by the molecular weight, Mi (in kg kmol−1 ). temperature as a function of water in the liquid (0–20 wt%) are
The molar flux of a compound through a permselective mem- shown in Fig. 5. The Antoine relationship was used to calculate
brane is a function of membrane transport properties and the par- psat
i
. The NRTL and Antoine parameters used herein are given in
tial vapor pressure driving force for transport, represented as:16 Tables S1 and S2, respectively, in the Supporting Information. The
activity coefficients for the solvents are in the narrow range of 1 to
Pi F 1.6 while those of water range more widely from 1 to almost 28.
ji = (p − pPi ) (5)
𝓁 i The total hypothetical feed-side vapor pressure for a PV liquid
feed is then the sum of the hypothetical partial pressures:
where Pi is the molar permeability of species i in the permselective
material (in kmol m m−2 s−1 kPa−1 ), 𝓁 is the thickness of the perms- ∑
pF∗ = xi 𝛾i psat
i
(8)
elective layer (in meters), and pFi and pPi are the partial pressures (in i
kilopascals) of species i in the feed-side fluid and permeate vapor,
respectively. Permeability is defined as the product of the solu- The hypothetical feed-side vapor mole fraction (xi∗ ) is defined as
bility and diffusivity of a mobile species in the dense permselec- the ratio of the hypothetical partial pressure to the hypothetical
tive membrane. For a prepared membrane with fixed or unknown total pressure (xi∗ = pF∗
i
∕pF∗ ). Even though VLE behavior is used
selective layer thickness, the ratio of Pi /𝓁, termed ‘permeance’ (Πi ) in the calculation, the hypothetical vapor is represented as a
498

(in kmol m−2 s−1 kPa−1 ), is often the most appropriate performance lowercase x to indicate it is a feed-side concentration.

wileyonlinelibrary.com/jctb Published 2019. This article is a U.S. Government work J Chem Technol Biotechnol 2020; 95: 495–512
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Review of pervaporation and vapor permeation process factors www.soci.org

L)
Figure 5. Activity coefficients of (a) water and (b) solvent for binary water/solvent mixtures as a function of water weight fraction in the liquid phase (ww
over the 0–20 wt% range. Values calculated with NRTL model at the normal boiling point temperature of each solvent. Values for methyl isobutyl ketone
(MIBK), and methyl tertiary butyl ether (MtBE) shown in the range of solubility of each with water.53,54

Figure 6. Effect of temperature on the partial vapor pressures of (a) water and (b) ethanol in equilibrium with a water/ethanol liquid mixture as a function
L ) over the 0–20 wt% range for temperatures ranging from 30 to 130 ∘ C. Values calculated with NRTL
of the water weight fraction in the liquid phase (ww
model.

Given this information about partial pressures, Eqn (5) can be rise. Operating at the highest practicable temperature will yield the
rewritten for V·P and PV operations as: highest feed-side partial pressures and, therefore, the highest par-
tial pressure driving force, resulting in the highest flux. Also evident
Pi from Fig. 6 is that the water partial pressure is strongly dependent
Vapor Permeation ∶ ji = (x pF − yi pP ) (9)
𝓁 i on composition while the ethanol partial pressure varies only mod-
estly with water concentration in this range. Consequently, as the
Pi P water content in the feed-side fluid decreases as the fluid traverses
Pervaporation ∶ ji = (x 𝛾 psat − yi pP ) = i (xi∗ pF∗ − yi pP ) the membrane grid or as a batch is processed, the feed-side vapor
𝓁 i i i 𝓁
(10) pressure of water will decrease. According to Eqn (5), for there to
The driving force for V·P is then controlled by altering the applied be a positive flux of water through the membrane, the permeate
feed-side and permeate pressures. In pervaporation, the applied partial pressure of water must be below the feed-side partial pres-
feed pressure plays little role, other than to maintain the fluid in sure of water. As indicated in Fig. 6, the feed-side partial pressure of
a liquid state. Instead, the feed-side temperature has the leading water may be quite small relative to that of the solvent and relative
role in PV driving force, primarily through the near-exponential to the initial water partial pressure, depending on the initial water
effect on saturated vapor pressure. The effects of composition and content. This may necessitate a lower absolute permeate pressure
temperature on the partial vapor pressures of water and ethanol for some, or all, of the system or operating time in order to motivate
in equilibrium with a liquid mixture in the range of 0 to 20 wt% water transport through the membrane.
water are shown in log–log plots in Fig. 6. Both water and ethanol The solvent in the mixture greatly impacts the partial vapor
partial pressures depend heavily on temperature at all concentra- pressure of water, even for the same water composition. First,
tions, increasing by a factor of 2 to 2.5 for every 20 ∘ C temperature
499

the type of solvent has an effect through the activity coefficient

J Chem Technol Biotechnol 2020; 95: 495–512 Published 2019. This article is a U.S. Government work wileyonlinelibrary.com/jctb
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org LM Vane

Figure 7. Partial vapor pressure of (a) water and (b) solvent in equilibrium with a water/solvent liquid mixture at 100 ∘ C as a function of the water weight
L ) over the 0–20 wt% range as calculated with NRTL model. Values for methyl isobutyl ketone (MIBK), and methyl tertiary
fraction in the liquid phase (ww
butyl ether (MtBE) shown in the range of solubility of each with water.53,54

as shown in Fig. 5. The y-axis intercept of each curve in Fig. 5(a) not usually practical in larger industrial systems. In cases where the
represents the infinite dilution activity coefficient for water (i.e. permeate pressures are non-negligible, the ratio of total feed vapor
𝛾w → 𝛾w∞ as xw → 0), which varies from just over 1 in DMF to pressure (pF for V·P or pF* for PV) to total permeate pressure pP ,
almost 28 in MtBE. For most of the solvents, 𝛾 w decreases markedly termed ‘the pressure ratio’ 𝜙, is an important design and operat-
as water content increases. However, the activity coefficients for ing factor, which can result in an actual observed separation quite
the solvents in the 0–20 wt% water range are close to 1, never different from the ideal membrane separation.
exceeding 1.6. The effect of solvent type and composition on The separation attributable to the membrane alone is the ratio
the partial vapor pressures of water and the solvent at fixed of the permeabilities (or permeances) of two compounds in the
temperature of 100 ∘ C are shown in Fig. 7. The same temperature membrane material, termed the molar permselectivity, 𝛼, or simply
was selected for all the mixtures to hold psat w
constant in this ‘selectivity’:
comparison and 100 ∘ C was selected because it is the normal P Π
𝛼 mem = 1 = 1 (11)
boiling point of water. In the 0–20 wt% water range, the partial P2 Π2
pressure of a solvent is nearly that of the pure solvent and, as
such, the solvent partial pressures at 100 ∘ C are ordered in the Combining Eqn (11) with Eqns (6), (9), and (10) results in the
figure based primarily on their normal boiling point. However, following expressions for V·P and PV, relating the separation fac-
water partial pressure varies with the solvent type based primarily tors calculated from observed values of x i and yi to the membrane
on the effect of the solvent on the activity coefficient of water. selectivity:
For mixtures containing 1 wt% water, the water partial pressure at
( )⎡1 − 1 y1
⎤ ⎡1 − 1 y1

100 ∘ C changes by over a factor of 20, from 3.1 kPa in methanol P1 𝜙 x1 𝜙 x1
Vapor Permeation ∶ 𝛽12
V·P
= ⎢ ⎥ = 𝛼 mem ⎢ ⎥
to 72.6 kPa in MtBE. Even within the four alcohols in the targeted P2 ⎢1 − 1 y2 ⎥ ⎢1 − 1 y2 ⎥
solvent list, the water partial pressure at 1 wt% water ranges from ⎣ 𝜙 x2 ⎦ ⎣ 𝜙 x2 ⎦

the 3.1 kPa in methanol to 17.2 kPa in 1-butanol. Thus, as solvent (12)
hydrophobicity increases, so does the water partial pressure in the ( )[ ] 1 y1
feed, increasing the PV driving force for permeation in Eqn (10). P1 (x1∗ ∕x1 ) ⎡ 1 − 𝜙 x1∗

Pervaporation ∶ 𝛽12
PV
= ⎢ ⎥
This will increase the driving force and flux for water removal from P2 (x2∗ ∕x2 ) ⎢ 1 − 1 y2 ⎥
⎣ 𝜙 x2∗ ⎦
the more hydrophobic compounds.
1 y1
⎡1 − 𝜙 x1∗

Effect of non-negligible permeate pressure on PV and V·P =𝛼 mem
𝛼12
VLE ⎢ ⎥ (13)
observations ⎢1 − 1 y2 ⎥
⎣ 𝜙 x2∗ ⎦
Equations (9) and (10) can be used to convert the flux and per-
meate composition results reported for a membrane under spec- According to these expressions, as 𝜙 increases, the right hand
ified operating conditions to membrane permeance and selectiv- bracketed terms become 1 and the separation factor for V·P
ity values. These membrane performance parameters can be used approaches the selectivity of the membrane (𝛽12 V·P
⇒ 𝛼 mem ), while
to better estimate performance under other conditions. In many the separation factor for PV approaches the product of the
studies reported in the literature, permeate partial pressures are membrane selectivity and the relative volatility of the feed-side
negligible compared to the feed partial pressures. This is due to PV
liquid (𝛽12 ⇒ 𝛼 mem · 𝛼12
VLE
). Thus, in PV, favorable VLE augments
the study of moderate feed water concentrations (5–15 wt%) and membrane selectivity while unfavorable VLE detracts from mem-
collection of the permeate using liquid nitrogen-cooled traps in brane selectivity. Fortunately, unfavorable VLE behavior can be
combination with vacuum pumps capable of delivering low abso- overcome with a high membrane selectivity to yield a separation.
500

lute pressures (< 0.1 kPa). Such permeate collection conditions are In this way, a water-selective membrane can remove water even

wileyonlinelibrary.com/jctb Published 2019. This article is a U.S. Government work J Chem Technol Biotechnol 2020; 95: 495–512
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Review of pervaporation and vapor permeation process factors www.soci.org

from the water/solvent systems in Fig. 4 exhibiting azeotropes or conditions, permeate conditions, and either the flow rate, if con-
low water/solvent relative volatilities. Ideally, the inherent proper- tinuous, or the charge volume and cycle time (𝜏) if a recirculating
ties of both the membrane and VLE work in the same enrichment batch operation. Unless the membrane area required can be met
direction. Since the separation in a V·P unit is not affected by VLE with a single membrane module, the membrane system will con-
behavior, there may be unfavorable VLE situations that would sist of a grid of membrane modules and housings arranged in par-
benefit from employing V·P instead of PV. allel and/or in series to accomplish the separation.
Minimizing membrane area is a general PV/V·P process design
Calculation of permeate composition objective because area represents both an initial capital cost and
an operating expense related to periodic replacement of the mem-
With regard to the permeate composition, Baker provides a very
brane elements. However, minimizing area may require feed-side
useful relationship for calculating the permeate concentration of
and permeate conditions that either result in large increases in
the higher permeance species in a binary gas or vapor mixture
other process costs or are just too extreme for the membrane or
from a specific feed-side concentration (x i ) and pressure ratio (eqn
module materials. The effect of membrane performance char-
8.19 in Baker16 ):
acteristics and several process parameters on membrane area,
( )⎡ permeate purity, and solvent recovery will be examined in this
𝜙 ⎢ 1 1 section. To do so, estimations of membrane performance will need
yi = x + +
2 ⎢ i 𝜙 to be established.
⎣ 𝛼 mem − 1

(
1 1
)2 4𝛼 mem xi ⎤⎥ Single-pass systems
− xi + + mem − mem (14) The most straightforward dehydration system to model is a
𝜙 𝛼 −1 (𝛼 − 1)𝜙 ⎥
⎦ single-pass operation in which a constant molar feed flow rate of a
solvent/water mixture (Ṅ F ) is treated with a membrane system to
Although developed for gas or vapor separation, Eqn (14) can reduce the water content from the inlet feed mole fraction (xwF ) to
be used for pervaporation by replacing x i with the hypothetical a final retentate water content (xwR ). If the illustrations in Fig. 1 are
feed-side vapor mole fraction (xi∗ ) and basing the pressure ratio taken as representing a differential membrane element (dA), then
on the total hypothetical feed-side pressure (pF* ). Clearly, both the differential loss of component i from the feed-side fluid stream
membrane selectivity and pressure ratio play a significant role (dṄ i ) as the fluid traverses the membrane element is equal to the
in determining the permeate composition. In the extremes of 𝜙 flux through the membrane multiplied by the membrane area as:
relative to 𝛼 mem , Eqn (14) simplifies to:
P
𝛼 mem xi
̇ i ) = −ji dA = − i (xi pF − yi pP ) dA
V · P ∶ dṄ i = d(Nx
if 𝜙 ≫ 𝛼 mem ("Selectivity − limited") ∶ yi = 𝓁
1 + xi (𝛼 mem − 1)
(15) P
̇ i ) = −ji dA = − i (x ∗ pF∗ − yi pP ) dA
PV ∶ dṄ i = d(Nx
𝓁 i
if 𝛼 mem ≫ 𝜙 ("Pressure Ratio − limited") ∶ yi = 𝜙xi (16) P
= − i (xi 𝛾i psat − yi pP ) dA (17)
Thus, under a pressure ratio-limited situation (Eqn (16)), the per- 𝓁 i

meate composition of the preferentially permeating compound Spreadsheets were created to estimate performance of
is independent of the membrane selectivity, dictated only by the single-pass V·P and PV systems, operating in a crossflow con-
applied pressure ratio and feed-side mole fraction – a concept that figuration as depicted in Fig. 1, by dividing the membrane unit
may be difficult to appreciate for those focused only on selectivity. into 100 sub-units linked in series, describing the transport of each
In such a case, raising the flux of the solvent will raise the water species in a sub-unit according to Eqn (17) (detailed in Supporting
flux since the ratio of fluxes is a constant (because yi is fixed). For Information). The soundness of using 100 sub-units was confirmed
example, under pressure ratio-limited conditions, switching to a by determining that an increase to 1000 sub-units resulted in rela-
membrane with lower selectivity but higher solvent permeance tive changes in calculated values of less than 1.4%. The advantage
will raise water flux while delivering the same permeate water of fewer sub-units is faster spreadsheet refreshing and iterative
composition. solving. The spreadsheets enable the study of different process
variables to assess their effect on system performance. A similar
spreadsheet was previously employed by the author’s group to
PV AND V·P PROCESS DESIGN estimate performance of a V·P system with multi-component
CONSIDERATIONS vapor feed streams.17,18 A rougher estimate of performance can
PV and V·P processes are primarily designed in one of two modes be made by applying several simplifying assumptions to Eqn (17).
of operation: single-pass or recirculating batch, as depicted in For a PV system, the assumptions are:
Figs 2 and 3. In single-pass mode, the membrane dehydration
• Temperature is constant (therefore psat values are constant)
system reduces the water content in a stream of constant flow i
• Activity coefficients are constant
rate from the inlet/feed mole fraction, xwF , to the desired prod-
• Total solvent flow rate is constant at the feed solvent flow rate,
uct/retentate solvent product mole fraction, xwR . In recirculating
Ṅ sF (i.e. negligible solvent permeates the membrane)
batch mode, the concentration of water in a discrete charge of
• Permeances are constant
water-contaminated solvent is reduced over time, 𝜏, from an initial
• Permeate pressure is negligible
mole fraction, xwF , to the desired final value for the product solvent,
0
xwF . The membrane area required for the dehydration task, A, is a
𝜏
With these assumptions and substituting Ṅ = Ṅ sF ∕(1 − xw ), Eqn
501

function of the membrane material and module, feed mixture, feed (17) (for PV) can be integrated to yield the following relationship

J Chem Technol Biotechnol 2020; 95: 495–512 Published 2019. This article is a U.S. Government work wileyonlinelibrary.com/jctb
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org LM Vane

between the retentate/outlet water composition (xwR ), feed/inlet Eqn (20) simplifies to (Pw ∕𝓁)xwt pF A for V·P and (Pw ∕𝓁)xwt 𝛾w psat w
A
water composition (xwF ), area, flow rate, membrane permeance, for PV. With these assumptions and substituting Nt = Ns0 ∕(1 − xwt ),
activity coefficient, and saturated vapor pressure: Eqn (20) for PV can be integrated with respect to time to yield the
[ R ] following relationship between the tank water composition at the
( )
xw∕ R)
(1−xw 1 1 A Pw cycle time 𝜏 (xw𝜏 ), initial tank water composition (xw0 ), area, initial
ln F + − ≅ − 𝛾w psat (18)
xw∕ F
(1−xw ) 1 − xw
R 1 − xw
F ̇
NsF 𝓁 w tank charge of solvent, membrane permeance, activity coefficient,
saturated vapor pressure, and processing time:
For small water concentrations, Eqn (18) simplifies further to: [ 𝜏 ] ( )
xw∕ 𝜏)
(1−xw 1 1 A𝜏 Pw
( ) ( ) ln 0 + − ≅− 0 𝛾w psat (21)
xwR A Pw xw∕ 0
(1−xw ) 1 − xw
𝜏 1 − xw
0 N s
𝓁 w
ln ≅ − 𝛾w psat (19)
xwF ̇
NsF 𝓁 w
Note that this equation is of the same form as Eqn (18) for a
For a single-pass V·P system with complete evaporation (see single-pass PV system, with the molar solvent flow rate Ṅ sF replaced
Option 2 in Fig. 3), Eqns (18) and (19) can be applied after substi- by the total moles of solvent in the initial tank charge divided by
tuting the total feed pressure pF for (𝛾w psat ). the processing time (Ns0 ∕𝜏). For small water concentrations, Eqn
w
(21) further simplifies to:
Recirculating batch systems ( 𝜏) ( )
x A𝜏 Pw
In a recirculating batch membrane process, the amount of each ln w0 ≅ − 0 𝛾w psat (22)
xw Ns 𝓁 w
species in the feed source tank (Nit ) changes with time as does
the concentration in the tank (xit ). In the membrane system, the which is of the same form as Eqn (19) for a single-pass system
concentration of each species changes with both time and posi- with the molar solvent flow rate Ṅ sF replaced by the total moles of
tion along the membrane system. The tank is charged initially solvent in the initial tank charge divided by the processing time
with amount of mixture N0 at water concentration xw0 . A stream is (Ns0 ∕𝜏).
removed from the source tank and processed through the mem- If a recirculating batch V·P system with complete evaporation
brane system, lowering the water content from the feed con- (Option 1 in Fig. 3) were being considered, Eqns (21) and (22) could
centration at a given time (xwF@t ) to the retentate concentration be used to estimate performance after substituting the total feed
at that time (xwR@t ), corresponding to the separation achieved by pressure pF for (𝛾w psat
w
), mirroring the situation for single-pass V·P
the membrane at that time. The retentate is returned to the feed with complete evaporation. If, however, the vapor for a recircu-
source tank. The source tank is considered well stirred such that lating batch V·P system is generated in a batch evaporator as in
the feed concentration to the membrane is the same as the tank Option 1 of Fig. S3 (where the liquid in a source vessel is heated
concentration at a given time (xwF@t = xwt ). The incremental change to generate a vapor through VLE in the vessel rather than a com-
in the amount of a compound in the source tank (dNit ) over a time plete evaporator), then Eqns (21) and (22) apply without alteration
increment (dt) is equal to the permeation through the membrane because of the VLE established in the batch evaporator.
system of area A at a particular time, represented as:
[ A ]
Pi IMPLICATIONS OF SIMPLIFIED PROCESS
V · P ∶ dNit = d(Nt xit ) = −ji A dt = − (xi pF − yi pP )dA dt
∫0 𝓁 @t EQUATIONS
[ ] While Eqns (18)/(19) and (21)/(22) are rough estimates of how a
A
Pi single-pass or recirculating batch process will reduce water in a sol-
PV ∶ dNit = d(Nt xit ) = −ji A dt = − (xi 𝛾i psat − y pP
)dA dt
∫0 𝓁 i i
@t
vent, they can be used to identify important factors that will affect
(20) performance of these systems. First, according to these equations,
where the left side of each relationship represents the change the area required to reduce water by a certain extent in a PV sys-
in moles of species i in the feed tank, ji is the area-averaged tem is proportional to the flow rate in a single-pass system or to the
flux at time t, and the integration on the right-hand side is over amount of material to be processed in a given time for a recirculat-
the area of the membrane system at the conditions of time t. ing batch operation and is inversely proportional to water perme-
A spreadsheet was created to estimate recirculating batch PV ance, water activity coefficient, and the saturated vapor pressure
system performance by dividing the batch cycle into 100 time of water. Because psat
w
increases rapidly with increasing tempera-
intervals, describing the transport of each species in a time interval ture, operating at the maximum temperature will greatly reduce
according to Eqn (20) (as described in Supporting Information). required membrane area or processing time. The corollary for V·P
A spreadsheet model of a recirculating batch V·P operation was systems is that operating at maximum pF will minimize area. Sec-
not developed because this type of operation is not particularly ond, for PV systems, the specific solvent in the mixture impacts
practical due to the large amount of energy required to evaporate water removal mainly through 𝛾 w . Since 𝛾 w can vary significantly
and condense the recirculation stream. from one solvent to another (as noted earlier, 𝛾w∞ ranges from
A rougher but simpler estimate of recirculating batch system about 1 to near 30 among the target solvents), the area or time
performance can be established by making the same assumptions required to accomplish a desired decrease in water concentration
as earlier for the single-pass system estimate, except assuming the can vary significantly between solvents. As indicated in Fig. 5 the
amount of solvent in the feed source tank is constant and equal to differences in 𝛾 w between solvents become less pronounced as
the amount of solvent in the initial tank charge (i.e. Nst = Ns0 ) rather water concentration increases. Third, the ratio of final and initial
than that the flow rate of solvent is constant. In addition, if it is water concentrations is roughly an exponential function of the
assumed that the single-pass removal of water in the membrane other parameters but is relatively independent of the initial con-
502

unit is small such that xwR@t ≅ xwF@t , then the right-side integral in centration. Thus, reducing water from 10 wt% to 1 wt% requires

wileyonlinelibrary.com/jctb Published 2019. This article is a U.S. Government work J Chem Technol Biotechnol 2020; 95: 495–512
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Review of pervaporation and vapor permeation process factors www.soci.org

approximately the same area as reducing water from 1 wt% to other, resulting in relatively small changes in permeance, partic-
0.1 wt%, despite a ten-fold difference in the amount of water ularly over small temperature changes (e.g. ΔT of 5 to 10 ∘ C). For
removed, based on the assumptions made. some materials, such as inorganic and amorphous perfluoro poly-
Just as illustrative are the conditions under which the simplifying meric materials, permeances are reasonably independent of water
assumptions used in establishing the simplified equations fail: concentration. However, hydrophilic polymers that plasticize upon
water sorption can exhibit large changes in permeance with water
Assumption: temperature is constant concentration. For example, the PV water permeance of a commer-
Since PV involves a phase change from the feed liquid to the cial poly(vinyl alcohol) membrane increased by 140% as water in
permeate vapor, the feed liquid cools as it traverses the membrane the water/ethanol feed increased from 0.7 to 15 wt%.19 The per-
system, thus temperature is not constant in PV and this will have a meance of such a material will depend on the water activity on
dramatic effect on local flux due to a drop in partial vapor pressure. both the feed and permeate sides of the selective layer. A variety
For example, based on the heat of evaporation of water and of models and methods of accommodating such fluctuations have
ethanol/water mixture heat capacities, reducing water content in been developed.20–22 An additional factor for V·P is superheat, in
ethanol from 10 wt% to 9 wt% requires 24.6 kJ of heat per kilogram which the feed-side vapor is heated above the dew point temper-
of the feed mixture and results in a 7.4 ∘ C temperature drop. If the ature to avoid condensation in the membrane module. Superheat
feed were at 100 ∘ C, this ΔT would result in a drop in the saturated reduces the activity of water and solvent experienced by the mem-
vapor pressure of water of 24%. As a result, heat may need to be brane that may alter the performance of the membrane.
added between, or in, membrane modules to counteract the heat
lost due to evaporation. V·P does not involve a phase change in Assumption: permeate pressure is negligible
the membrane module, so this assumption is generally reasonable. This assumption is often the first to be made because it greatly
The corollary assumption for V·P is that the total feed-side pressure simplifies flux equations but is the hardest to justify in commer-
is constant. However, in real systems, pF will fall due to viscous cial PV or V·P processes due to the cost of supplying low abso-
flow pressure drop as the vapor moves through the membrane lute vacuum pressures. In addition, if low retentate water concen-
module(s). trations are required, low feed-side partial pressures of water will
result – making it difficult to achieve permeate pressures that are
Assumption: activity coefficients are constant negligible by comparison. Further complicating this assumption
As illustrated in Fig. 5 and discussed earlier, 𝛾 w can vary signifi- is the permeate pressure drop contributed by the porous support
cantly with water content and this variability is solvent-dependent. layers that provide mechanical stability to the thin, permselective
Thus, this assumption is invalid when: 𝛾 w varies significantly with layer as well as the pressure drop due to vapor flow through the
water content, the initial water content is high, and the change membrane element.
in water content is large. As shown in Fig. 5, reducing water from Some of these non-idealities can be partially accounted for in
20 to 0.1 wt% results in 𝛾 w increasing by at least 50% for all of the the simplified equations. The effect of temperature on psat and of
w
solvents except methanol and DMF. However, as water is reduced concentration on 𝛾 w can be approximated in the PV performance
from only 1 to 0.1 wt%, 𝛾 w increases by 50% or more for just one estimations using the geometric mean of the two parameters
solvent, MtBE. between the feed and retentate conditions, akin to the use of
VLE
geometric mean of relative volatility (𝛼 12 ) in determining the
Assumption: solvent molar flow rate or amount of solvent minimum number of distillation stages in Eqn (3). Similarly, an
in the feed tank is constant average permeance could be employed to account for the effect
This assumption is reasonable when the amount of solvent in of concentration on permeance for some membranes. Of the
the permeate is low relative to that in the feed, typically when earlier assumptions, the spreadsheet calculations used herein only
membrane selectivity is high and permeate pressure is low – or if require constant permeances within a membrane unit. Inclusion of
membrane area is low as might be the case for a small (e.g. < 50%) permeances as functions of temperature and component activity
relative change in water concentration. For all of the solvents, a are being considered for future work.
selectivity of 100 or more will result in less than 10% transfer of
solvent to the permeate for a reduction in water content from
10 to 0.1 wt%, when permeate pressure is negligible. The effect IMPACT OF MEMBRANE PROPERTIES AND
of selectivity and permeate pressure on solvent transfer to the PROCESS VARIABLES ON SYSTEM DESIGN
permeate will be explored later. In this section, the effect of process parameters on system per-
formance will be evaluated using the estimating equations devel-
Assumption: permeances are constant oped earlier and the more in-depth spreadsheet calculations
In theory, permeances are less dependent on temperature and developed from Eqns (17) and (20) for single-pass PV, recirculat-
concentration than fluxes because they have been normalized by ing batch PV, and single-pass V·P (see Supporting Information for
partial pressures – which are highly dependent on temperature descriptions of the spreadsheet calculations). The effect of mem-
and concentration. However, permeance values reported in the lit- brane properties on ethanol/water separation performance will be
erature or in vendor datasheets were obtained under experimental studied based on the permeance and selectivity ranges outlined
conditions that might not relate directly to the conditions of the in the materials review paper for commercially available mem-
solvent drying process being considered. Permeability is the prod- brane materials. The benchmark separation discussed later will
uct of solubility and diffusivity. As a result, parameters that alter involve reducing water in ethanol from 10 to 0.5 wt% (22.1 to
solubility and diffusivity will affect permeability and permeance. 1.27 mol%) using a membrane with a constant water permeance
For example, increasing temperature generally decreases solubil- of 2000 GPU and a constant water/ethanol selectivity of 2000. This
503

ity but increases diffusivity. These changes partially offset each set of conditions was selected to be representative of the middle

J Chem Technol Biotechnol 2020; 95: 495–512 Published 2019. This article is a U.S. Government work wileyonlinelibrary.com/jctb
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org LM Vane

ground of the seven common PV/V·P membrane materials remov-


ing water from ethanol/water mixtures by pervaporation for feeds
containing 5–15 wt% water at temperatures of 50 to 80 ∘ C that was
detailed in the previous paper.9 The 10 to 0.5 wt% water bench-
mark separation was selected to illustrate the removal of water
from a concentration above the azeotrope (4 wt% water in the
ethanol azeotrope) to a product concentration that is consistent
with fuel ethanol specifications (0.3 wt% in the European Union23
to 1.3 wt% in the United States24 ). The benchmark product water
level of 0.5 wt% is also consistent with generally acceptable water
impurity levels in recovered solvents to retain solvency (1% water)
or to avoid affecting reaction equilibria (0.1% water).25 Product
water levels of 100 ppm (0.01%) or lower might be necessary if
water deactivates a catalyst or reagent in the intended applica-
tion of the reclaimed solvent. For other solvents, the target prod-
uct water concentrations will likely be similar to those in ethanol,
although the initial water concentration will depend on a number
Figure 8. Membrane area of a single-pass pervaporation (PV) system
of factors, including the azeotrope for the specific solvent/water required to reduce the water concentration of an ethanol/water solution
system. The effect of varying the feed and product water con- from 10 wt% in the feed stream to 0.1, 0.5, 1, or 5 wt% as a function of the
centrations on ethanol/water separation system design will be total permeate pressure. Values calculated with single-pass PV spreadsheet
explored. under benchmark conditions: isothermal operation at 70 ∘ C, 0.5 kg s−1
ethanol feed rate, 2000 GPU water permeance, and water/ethanol 𝛼
The benchmark membrane system is a 70 ∘ C isothermal,
= 2000. The closed symbols on the y-axis are the areas calculated from sim-
single-pass PV process with a solvent feed rate of 0.05 kg s−1 plified Eqn (18) with geometric mean of 𝛾w psat w . Open symbols represent
(about two million liters per year) operating with negligible per- the permeate pressure at which the area has increased 50% from the area
meate pressure (pP = 0). For a benchmark recirculating batch calculated at pP = 0 for a specific retentate concentration. Arrows along
PV system, an initial solvent charge of 1440 kg, equal to 8 h of a the x-axis are estimates of the maximum practical permeate pressure as
described in the text.
0.05 kg s−1 solvent flow, is assumed to be processed in an 8 h cycle,
all other parameters being the same as the single-pass benchmark
conditions. symbols at a 50% area increase. For instance, the benchmark sys-
tem of 0.5 wt% retentate needs 50% more area at pP =0.86 kPa than
at pP =0, while a system treating down to a higher retentate con-
Retentate concentration and permeate pressure centration of 5 wt% water would see a 50% increase at a higher
As calculated with the single-pass PV spreadsheet, the membrane permeate pressure of pP =3.6 kPa. As mentioned earlier, the per-
area required to perform the benchmark separation is 75.5 m2 . meate vacuum is usually generated with a condenser in combina-
The estimate of area for this separation from simplified Eqn (18) tion with a vacuum pump. Lower permeate pressures require lower
with a geometric mean activity coefficient is 78.2 m2 , which is only condenser temperatures or booster vacuum pumps, each adding
3.5% higher than that from the spreadsheet. The impact of varying capital and operating costs. Condensing pure water at absolute
retentate concentration and permeate pressure on the membrane pressures of 0.86 and 3.6 kPa requires temperatures of 5 and 27 ∘ C,
area required for the separation, keeping all other parameters the respectively. Simply considering this 22 ∘ C difference in condenser
same as the benchmark scenario, is displayed in Fig. 8. Each curve temperature makes the impact of permeate pressure on perme-
represents a different retentate concentration of water, ranging ate condensation conditions apparent. The presence of solvent in
from 0.1 to 5 wt%, with total permeate pressure as the x-axis. the permeate reduces condenser temperatures further due to the
The intercept of each curve with the y-axis is the membrane area impact of solvent on the bubble point of the permeate mixture.
needed when the permeate pressure is negligible (e.g. 75.5 m2 for Figure S8 in the Supporting Information displays the effect of vary-
0.5 wt% retentate). The closed symbols shown on the y-axis are the ing both feed and retentate water concentration in the range of
rough area estimates calculated from Eqn (18), for which pP =0 had 0.1 to 20 wt% on membrane area, permeate solvent concentration,
been assumed in the derivation. Each open symbol on the curves and evaporation energy of the, otherwise, benchmark single-pass
represents the point at which the area has increased 50% from the PV system.
area calculated at pP =0 for a specific retentate concentration. A The benchmark recirculating batch PV system should, in the-
50% change in membrane area will be used herein as a subjective ory, require the same area as the benchmark single-pass system
indicator that a variable has reached a value that causes a notable because the amount of the batch solvent charge and cycle time
change in membrane area. were chosen to parallel the single-pass solvent flow rate. In fact,
It is evident from the proximity of the solid symbols to the y-axis at 78.2 m2 , the estimate of area for the recirculating batch PV
intercepts of the curves that the areas calculated from simpli- operation from Eqn (21) is exactly the same as the estimate for
fied Eqn (18) are reasonable estimates of the spreadsheet calcu- the single-pass PV operation from Eqn (18). The fuller spread-
lations for pP =0, with relative differences ranging from only 0.2 sheet calculation for the recirculating batch system yields an
to 5.4%. For non-negligible permeate pressures, as expected, the area of 74.4 m2 , which is only slightly lower than the single-pass
required area increases as pP increases due to the reduced driv- spreadsheet result of 75.5 m2 . Thus, the recirculating batch and
ing force for flux through the membrane. However, the magnitude single-pass benchmark systems yield analogous results for equiv-
of this impact depends on the specified retentate water concen- alent operating parameters. In addition, the permeate pressure
tration – as indicated in Fig. 8 by the pressure ranges where the that yields a 50% increase in recirculating batch system area is
504

areas rapidly increase and by the relative positions of the open 0.87 kPa, comparable to the 0.86 kPa observed for the single-pass

wileyonlinelibrary.com/jctb Published 2019. This article is a U.S. Government work J Chem Technol Biotechnol 2020; 95: 495–512
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Review of pervaporation and vapor permeation process factors www.soci.org

system. Thus, the spreadsheet calculations yield similar area results where x w for a single-pass system is either xwR for V·P or xwR∗ for PV.
under the benchmark parameters even when permeate pressure is For recirculating batch systems, the end-of-cycle tank concentra-
varied. tions of xw𝜏 (V·P) or xw𝜏∗ (PV) would be used. The maximum total
A single-pass V·P system in which the 10 wt% water feed stream permeate pressure would then be the total feed-side vapor pres-
is fully evaporated at 70 ∘ C, generating a total feed pressure of sure under the retentate, or final, conditions (pR ) divided by 𝜙min
71.6 kPa (based on a dew point calculation with CHEMCAD), would as:
require 76.3 m2 of membrane, very similar to that of the PV system. pPmax ≅ pR∕𝜙min (24)
A 50% increase in V·P membrane area is estimated to occur at a
permeate pressure of 0.79 kPa, also similar to that of the PV system. For processes involving only small reductions in water concen-
The V·P area is slightly different than that of the PV system because tration, the feed and retentate partial pressures of water (pFw and
of the difference between the mole fraction and partial pressure of pRw ) may not be very far apart. In such cases, pPmax based only on
water from the complete evaporator in the V·P system and those retentate conditions using the earlier calculation would apprecia-
corresponding to the hypothetical vapor in equilibrium with the bly reduce the water vapor pressure driving force not just at the
liquid and the membrane in a PV system. end of the system or cycle, but throughout the membrane system,
Individual pervaporation membrane modules range in active leading to a large increase in membrane area. In those situations,
membrane area from 1 to 50 m2 , depending on the type of mem- an alternate pPmax could be calculated based on a specified allow-
brane. Plate-and-frame modules, in which flat sheets of membrane able relative increase in membrane area (Ψ). For example, the 50%
are layered with feed and permeate spacers, have been used most increase in membrane area used as a reference in Fig. 8 would cor-
extensively for pervaporation dehydration and have reported respond to Ψ = 1.5. Using the geometric mean of the feed-side
active areas of up to 50 m2 .16,26,27 By comparison, a 1 m long, 4 in. water partial pressure, pPmax would be estimated in this way as:
(10 cm) diameter spiral wound membrane module contains 3 √
to 6 m2 of active membrane and an 8 in. diameter spiral wound ( ) pFw pRw
1
module contains from 20 to 40 m2 .16 Individual zeolite mem- pPmax ≅ 1 − (25)
brane tubes of 12 mm outer diameter and 0.8 m length have an
Ψ ywgm
active area of 0.03 m2 (based on outer active surface). Multi-tube
where ygm
w is the maximum possible water concentration in the
zeolite modules with areas up to 7.3 m2 have been described
permeate associated with the geometric mean
√of the feed and
for pervaporation dehydration.28,29 Multi-element hybrid silica
membrane modules with areas of up to 3.8 m2 are also available.30 retentate hypothetical feed concentrations (i.e. xwF∗ xwR∗ ), as calcu-
Similarly, multiple spiral wound or plate-and-frame modules can lated with Eqn (15) for the ‘selectivity limited’ condition.
be plumbed together in a single pressure vessel, allowing for The lesser of the permeate pressure estimates from Eqns (24)
much larger membrane areas in a single assembly. For example, and (25) would then be used as pPmax . The lesser values are pre-
an early industrial pervaporation plant drying ethanol from 7 wt% sented as arrows along the x-axis of Fig. 8 in order of increas-
water to 0.2 wt% water consisted of 42 plate-and-frame modules, ing water in the retentate (color coded to match the respective
divided between three pressure vessel assemblies. Each assembly curve). Calculations of pPmax are provided in Table S5 of the Sup-
contained about 50 m2 of membrane, for an average assembly porting Information. For example, the feed and retentate partial
area of 700 m2 and total system area of about 2100 m2 .27,31 Thus, pressures of water in the benchmark single-pass PV scenario are
the calculated membrane areas presented in Fig. 8 are well within 14.0 and 0.98 kPa, respectively, with xwF∗ =0.193, xwR∗ =0.0134, and
the practical range of commercial membrane module/assembly pR =73.2 kPa. From Eqn (15), the maximum permeate composition
designs. The exact design of the membrane system will be deter- associated with the geometric mean feed-side concentration (ywgm )
mined by process specifics, such as reheating requirements, and is 0.991. Based on Eqns (23) and (24), 𝜙min =72 and pPmax =1.0 kPa.
membrane/module type. Alternatively, with Ψ = 1.5, Eqn (25) yields pPmax =1.2 kPa. Thus, the
lesser of the two is 1.0 kPa (the value indicated in Fig. 8), which is
similar to the 0.87 kPa permeate pressure associated with a 50%
Estimation of maximum practical permeate pressure area increase based on the spreadsheet calculations. For the 0.1
Permeate pressure clearly plays a significant role in determining and 0.5 wt% retentate examples, Eqn (24) yielded the lesser value
many PV and V·P system characteristics. It would be useful to of pPmax , while pPmax from Eqn (25) was the lesser value for the
have an estimate of the maximum practical permeate pressure 1 wt% and 5 wt% retentate examples. For each curve, pPmax is sit-
(pPmax ) and maximum permeate condenser temperature for a given uated in the region where the required area begins to increase
set of feed conditions. One such estimate can be derived from sharply with increasing total permeate pressure. As a result, the
the target concentration of water in the solvent product and the estimate of pPmax is useful to assess when total permeate pres-
operating temperature. According to the flux equations, the partial sure will impact process performance and the area required. In
pressure of water in the permeate must be below that on the feed addition, the maximum permeate condenser temperature can be
side of the membrane to maintain a flux of water (i.e. pPw < pFw estimated from the bubble point temperature of the permeate
in Eqn (5)). The lowest feed-side partial pressure of water occurs at pPmax .
at the lowest water concentration – either in the retentate of a An additional limitation to achieving low permeate pressures
single-pass system or in the recirculation tank at the end of a batch via condensation is the freezing temperature of the permeate,
cycle. A rough estimate of the minimum pressure ratio can be which limits both the minimum temperature achievable in a
calculated by dividing the maximum permeate composition from simple liquid condenser and the associated vapor pressure at
Eqn (15) by the specified retentate water concentration, yielding: that temperature.32,33 For example, permeates with low solvent
concentrations will freeze near the melting point of water. Pure
𝛼 mem water ice at 0 ∘ C has a water vapor pressure of 0.6 kPa. Even ice
𝜙min ≅ (23) subliming at −20 ∘ C exerts a vapor pressure of 0.1 kPa.34
505

1 + xw (𝛼 mem − 1)

J Chem Technol Biotechnol 2020; 95: 495–512 Published 2019. This article is a U.S. Government work wileyonlinelibrary.com/jctb
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org LM Vane

Figure 9. The effect of water/ethanol selectivity of membrane (𝛼 = 20, 200, 2000, 20 000) in single-pass pervaporation (PV) system on (a) membrane area;
(b) bubble point temperature of the permeate; (c) concentration of ethanol in the permeate; (d) fraction of ethanol recovered in the retentate product. All
are shown as a function of the total permeate pressure. Curves in (b) only shown above the freezing point temperature. Values calculated with single-pass
PV spreadsheet under the following benchmark conditions: 10 wt% water in feed, 0.5 wt% water in retentate, isothermal operation at 70 ∘ C, 0.5 kg s−1
ethanol feed rate, and 2000 GPU water permeance.

Water permeance and selectivity selectivity values of 20, 200, 2000, and 20 000 on (a) membrane
The review of membrane materials noted the wide range of area; (b) bubble point temperature of the permeate (i.e. permeate
water permeances and water/ethanol permselectivities reported condenser temperature); (c) concentration of ethanol in the per-
for commercially available solvent dehydration membranes.9 To meate; (d) fraction of ethanol recovered in the retentate product
investigate the effect of changing water permeance and/or selec- are shown in separate graphs in Fig. 9. The x-axis variable in all
tivity on the performance of the benchmark single-pass PV system, graphs in Fig. 9 is the total permeate pressure. All other benchmark
the spreadsheet PV calculations were carried out for water per- conditions are retained. The benchmark case, with a selectivity
meances and selectivities ranging from 500 to 20 000 GPU and of 2000, is shown as a solid blue line in the graphs. Figure 9(a)
20 to 20 000, respectively. Varying water permeance while hold- confirms that the membrane area required when the permeate
ing selectivity constant resulted only in a change in the membrane pressure is negligible is independent of selectivity – as indicated
area required to reach the specified retentate water concentration. by the convergence of the curves at the y-axis. However, as total
Area was proportional to the inverse of the water permeance, as permeate pressure increases, the impact of selectivity becomes
is also predicted by the simplified performance equations (Eqns significant as indicated by the diverging curves. The higher the
(18)/(19) and (21)/(22)). Modeling the system with a non-negligible selectivity, the more sensitive membrane area is to permeate
permeate pressure did not alter this inverse relationship. The prod- pressure. The reason for this behavior is that a lower selectivity
uct A (Pw /𝓁) was constant for a fixed permeate pressure. All other results in ethanol diluting the water in the permeate, effectively
calculated process values, such as the concentration of ethanol reducing the partial pressure of water in the permeate. As shown
in the permeate or the permeate pressure that resulted in a 50% in Fig. 9(c), there is a dramatic difference in the concentration of
increase in required area compared to a negligible permeate pres- ethanol in the permeate for each decade change in selectivity.
sure, were unaffected by a change in water permeance at constant When pP = 0, the overall permeate ethanol concentration covers
selectivity. the range of 0.15 wt% for 𝛼 = 20 000 to 58.8 wt% when 𝛼 = 20.
Altering water/ethanol selectivity has the opposite effect, it More importantly, the local permeate concentration of ethanol at
506

changes almost all calculated process values. The effects of the retentate end of the system with pP = 0 is 90.9 wt% (80 mol%)

wileyonlinelibrary.com/jctb Published 2019. This article is a U.S. Government work J Chem Technol Biotechnol 2020; 95: 495–512
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Review of pervaporation and vapor permeation process factors www.soci.org

for 𝛼 = 20 but only 0.99 wt% (0.39 mol%) for 𝛼 = 20 000. As a result,
the low selectivity membrane inherently dilutes the permeating
water with ethanol. Whereas the benchmark system (𝛼 = 2000)
requires 50% more area at a permeate pressure of 0.86 kPa, the
system with 𝛼 = 20 does not increase in area by 50% until a much
higher permeate pressure, 3.4 kPa. Higher permeate pressures also
result in more favorable (i.e. higher) condensation temperatures,
presented in Fig. 9(b) as the bubble point of the permeate compo-
sitions in Fig. 9(c) for the range of membrane selectivities (bubble
point calculated with CHEMCAD). For a given permeate pressure,
a higher permeate ethanol concentration necessitates a lower
temperature to completely condense the vapor. Higher permeate
pressure results in a higher permeate ethanol concentration, but
the bubble point temperature is higher due to the higher partial
pressures.
The bubble point curves are only shown for temperatures above
the freezing points of the corresponding permeate compositions
shown in Fig. 9(c).32,33,35,36 Except for the lowest membrane selec- Figure 10. Effect of pervaporation (PV) temperature on the membrane
area required in a single-pass PV system to reduce the water concen-
tivity, permeate pressures achievable by liquid condensation will tration of an ethanol/water stream from 10 wt% to 0.5 wt% as a func-
be above 0.7 kPa. As illustrated in Figs 8 and 9(a), this permeate tion of the total permeate pressure. Values calculated with single-pass
pressure may result in a sizeable increase in membrane area. In PV spreadsheet under the following benchmark conditions: isothermal
order to achieve lower permeate pressures, a more complex per- operation, 0.5 kg s−1 ethanol feed rate, 2000 GPU water permeance, and
water/ethanol 𝛼 = 2000. The closed symbols on the y-axis are the areas
meate processing system is required, such as inserting a booster calculated from simplified Eqn (18) with geometric mean of 𝛾w psat w . Open
vacuum pump before the condenser, designing a permeate freez- symbols represent the permeate pressure at which the area has increased
ing/thawing system, or adding an absorption/desorption unit. For 50% from the area calculated at pP = 0 for a specific temperature.
the ethanol/water system, a lithium bromide absorber has been
proposed, utilizing the ability of this salt to reduce the vapor pres-
sure of both ethanol and water.32,33,37,38 Effect of feed temperature on PV system performance
Although the higher permeate pressures possible with lower Because the partial pressures of water and ethanol increase dra-
selectivity membranes have the potential to lower the capital and matically with increasing temperature (see Fig. 6), the membrane
operating costs associated with the vacuum system, they also area required to achieve a given reduction in water concentration
result in an unwelcome reduction in the fraction of ethanol recov- with a PV system will vary significantly with feed temperature. This
ered in the retentate product, as shown in Fig. 9(d). The maxi- effect on membrane area is exhibited in Fig. 10 for a single-pass
mum ethanol recovery for 𝛼 = 20 is 85% whereas the benchmark 𝛼 PV system operated at temperatures ranging from 30 to 110 ∘ C,
= 2000 system has a maximum solvent recovery of 99.8%. Further, all other properties being those of the hypothetical benchmark
the resulting higher level of solvent in the permeate may neces- single-pass PV system. Each curve represents areas calculated for a
sitate more advanced and costly processing, or disposal options, fixed feed temperature with the area plotted as a function of total
for the permeate. A fractional condenser could be added to the permeate pressure to illustrate the interplay between feed tem-
permeate condenser system to purify the permeate vapor. Such perature and permeate pressure. As in Fig. 8, the intercept of each
a ‘dephlegmator’ condenser has been demonstrated for separat- curve with the y-axis is the area corresponding to negligible per-
ing ethanol/water permeate vapors from a PV system.39,40 Alterna- meate pressure (e.g. A = 75.5 m2 for the benchmark 70 ∘ C feed) and
tively, the permeate condenser/vacuum system could be divided the closed symbols shown on the y-axis are the areas calculated
into two or more zones, each capturing the permeate vapor from from Eqn (18). As with the effect of varying retentate concentra-
a section of the membrane system. The pressure and condenser tion on membrane area, the Eqn (18) calculation provides a reason-
temperature of each zone could then be optimized with the poten- able estimate of the spreadsheet calculations for membrane area
tial to process the permeate condensate from each membrane at pP = 0. Each open symbol represents the permeate pressure at
section separately.41 A more complex system might involve utiliz- which the calculated area has increased 50% from the area calcu-
ing membranes with different properties in each section to further lated at pP =0 for a specific feed temperature.
optimize the separation and permeate processing.42 Finally, if the The vertical spacing of the curves and the log scale of the area
solvent recovery system already consists of a batch distillation or axis in Fig. 10 indicates that temperature does have a marked
evaporation unit, then that unit could be used to strip the sol- effect on the area required for a PV separation. The minimum area
vent from batches of condensed permeate containing unaccept- at 110 ∘ C is calculated to be 17 m2 whereas it is 534 m2 at 30
ably high concentrations of solvent. ∘ C. This 31-fold increase explains the motivation to operate PV
These same permeate management considerations apply to systems at the highest practical temperature. Despite this dramatic
single-pass V·P systems. For recirculating batch PV systems, the change in required membrane area, other process parameters
permeate condensate from different time intervals of the batch are unchanged, or only modestly altered, with temperature. For
cycle could be collected and managed separately, according to example, the concentration of ethanol in the permeate (pP = 0) for
the solvent concentration. Additionally, a booster vacuum pump the 30 ∘ C feed is calculated to be 1.61 wt%, but is calculated to
upstream of the permeate condenser could be brought on-line change only to 1.37 wt% for a 110 ∘ C feed.
at a certain time in the batch cycle to reduce the permeate As pP increases at each temperature in Fig. 10, the predicted
pressure and boost water partial pressure driving force to reduce membrane area increases due to the reduced driving force for flux
507

membrane area, cycle time, and permeate solvent concentration. through the membrane. However, the magnitude of this impact

J Chem Technol Biotechnol 2020; 95: 495–512 Published 2019. This article is a U.S. Government work wileyonlinelibrary.com/jctb
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org LM Vane

depends on the specified feed temperature as indicated in Fig. 10


by the relative positions of the open symbols that note a 50% area
increase for each curve. Systems operating at higher temperatures
are able to accommodate higher permeate pressures before sys-
tem area is impacted. As before, the benchmark system operat-
ing at 70 ∘ C needs 50% more area at pP =0.86 kPa than at pP =0,
while systems operating at 30 and 110 ∘ C are predicted to need
50% more area at pP of 0.12 and 3.8 kPa, respectively. The high
membrane area and low permeate pressure projected for a 30 ∘ C
PV operation make such low feed temperatures impractical unless
thermally-sensitive materials are present.
Increasing temperature affects the performance of recirculating
batch PV systems in the same manner as single-pass PV systems,
dramatically reducing the membrane area required to reduce the
water in a given initial charge of water/ethanol mixture over a fixed
cycle time. For single-pass V·P operations, the effect of tempera-
ture is through the partial pressures and total feed pressure gen- Figure 11. Effect of the reheat temperature drop threshold of a single-pass
erated in a complete evaporator operated at the designated tem- pervaporation (PV) system (solid lines) and the maximum temperature
perature. Since those pressures are determined by the evaporation drop in a batch PV system (dashed lines) on the membrane area required
to perform the benchmark separation for systems operating with negligi-
temperature, the effect of temperature on V·P system performance ble permeate pressure and with a permeate pressure that is predicted to
and membrane area will mirror that of the single-pass PV system. cause a 50% increase in area for an isothermal system. Values calculated
with single-pass and batch PV spreadsheets under the following bench-
mark conditions: 10 wt% water in feed, 0.5 wt% water in retentate, feed
Effect of temperature drop on PV system performance
temperature 70 ∘ C, 0.5 kg s−1 ethanol feed rate or 1440 kg ethanol batch
The performance calculations described earlier assumed isother- charge, 2000 GPU water permeance, and water/ethanol 𝛼 = 2000.
mal operation of the PV systems. Unfortunately, as noted in the
discussion of assumptions, PV involves a phase change from the
required membrane area increases relatively linearly and modestly
feed-side liquid to the permeate vapor resulting in a cooling of the
relative to both reheat threshold and ΔT max . A 50% increase in
feed-side liquid. This temperature change is small if a small amount
required membrane area for pP = 0 is predicted to result from
of permeate is generated per unit feed liquid, as might be the case
a 16 ∘ C inter-stage reheat threshold or batch ΔT max of 21 ∘ C.
for a small change in water concentration. However, the enthalpy
Consequently, allowing both single-pass and recirculating batch
of vaporization (ΔHivap ) of each species considered herein is large
systems to operate in a temperature window between the feed
compared to the heat required to alter the temperature of a sol-
temperature of 70 ∘ C down to about 50 ∘ C would only increase
vent/water liquid mixture. As a result, it does not require much
area by 50%.
water evaporation to cause a noticeable temperature drop and, as
Introducing a non-negligible permeate pressure changes the
noted earlier, system performance is strongly dependent on tem-
dynamic of temperature drop in a single-pass system, but not
perature.
for a recirculating batch system. As indicated in Fig. 11, the area
If the temperature drop is physically impractical, or simply more
required for a single-pass system rises rapidly with reheat thresh-
than desired, then heat must be added. In a single-pass system,
old when pP = 0.86 kPa, with a 50% increase above the base 113 m2
this usually involves inter-stage heat exchangers located so the
predicted to occur at a reheat threshold of only 7 ∘ C, noticeably
temperature change does not exceed a target threshold.43 This
lower than the 16 ∘ C threshold that caused a 50% area increase
involves a tradeoff as a smaller reheat threshold will necessitate
when pP = 0 was assumed. However, for a recirculating batch sys-
more, but lower power reheaters whereas a larger reheat threshold
tem operated with pP = 0.87 kPa, a 50% increase above the base
will reduce the number of reheaters but cause portions of the
112 m2 is predicted to occur when ΔT max is 21 ∘ C, the same as for
membrane system to operate at lower temperatures, depressing
pP = 0. The reason for the difference between recirculating batch
system performance and increasing membrane area due to a
and single-pass systems at non-negligible permeate pressures is
lower average operating temperature.44 In a recirculating batch
that the fixed reheat threshold for single-pass PV results in the
PV system, inter-stage reheaters could be used, but it could also
lowest temperature coinciding with the lowest water concentra-
be operated with just a feed heater and a higher recirculation
tion, severely lowering the feed-side vapor pressure of water. Con-
rate – in which case the maximum temperature drop (ΔT max ) for
versely, with the recirculating batch system, ΔT max is observed in
the system over the cycle time could be specified, which would
the initial time interval and diminishes as the cycle proceeds due
then specify a minimum recirculation rate given the separation
to less water evaporating per unit circulating fluid near the target
required in that time interval. The maximum temperature drop in
final water concentration. The impact of non-negligible permeate
a recirculating batch cycle is expected to be observed at the outset
pressure on single-pass reheat threshold could be moderated by
of the batch cycle due to the relatively high water flux at the initial
allowing a higher reheat threshold in the first membrane sections
water concentration in the cycle.
when the water is highest while reducing the reheat threshold in
The effects of single-pass inter-stage reheat threshold and batch
the last membrane sections, or by simply adding an inter-stage
ΔT max on membrane area and system performance, as calculated
heater later in the flow path.
with the spreadsheets based on Eqns (17) and (20), are shown
in Fig. 11 for two permeate pressure conditions: the benchmark
case of pP = 0 and the permeate pressure resulting in a 50% Effect of feed-side pressure drop on V·P system performance
increase in area for isothermal operation (0.86 or 0.87 kPa). For the As noted previously, the controlling parameter for V·P system
508

curves representing pP = 0 (the lower two curves in Fig. 11), the performance is the applied feed-side pressure. Because there is no

wileyonlinelibrary.com/jctb Published 2019. This article is a U.S. Government work J Chem Technol Biotechnol 2020; 95: 495–512
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Review of pervaporation and vapor permeation process factors www.soci.org

phase change in V·P, temperature drop due to permeation is not


a decisive issue for V·P as it is in PV. However, feed-side pressure
drop is a critical issue. The simple act of the feed vapor flowing
through the membrane module causes a viscous flow pressure
drop, the result of which is similar to that of temperature drop in
PV – a decline in the feed-side partial pressure of both species and
a drop in the driving force. The net result is an increase in required
membrane area and/or a need to operate with a lower permeate
pressure. The single-pass V·P spreadsheet allows for consideration
of a feed-side pressure drop as a fixed pressure drop for each
sub-unit of the system.

Effect of solvent on general system performance


The previous sections investigated how changes in process vari-
ables and membrane performance characteristics are predicted
to affect membrane area requirements, permeate purity, sol-
vent recovery, and permeate condenser/vacuum demands for the
removal of water from ethanol. In this section, the impact of chang-
ing the solvent in the mixture from ethanol to one of the other Figure 12. Effect of solvent in a binary water/solvent mixture on the area
targeted solvents will be assessed. In Fig. 7, the impact of solvent required for water reduction for single-pass pervaporation (PV) operating
type on the partial vapor pressures of both water and of the sol- at 100 ∘ C as a function of total permeate pressure. All areas are relative to
vent at 100 ∘ C were presented for water content up to 20 wt%. The the area required when ethanol is the solvent and pP =0 (Amin = 23.9 m2 ).
Values calculated with single-pass PV spreadsheet under the following
more hydrophobic solvents exhibited higher water partial pres- benchmark conditions: 10 wt% water in feed, 0.5 wt% water in retentate,
sures. How these partial pressure differences translate to PV system isothermal operation, 0.5 kg s−1 solvent feed rate, 2000 GPU water per-
performance was assessed first by calculating the areas required meance, and water/solvent 𝛼 = 2000. Methyl isobutyl ketone (MIBK), and
to reduce the water content from 10 to 0.5 wt% for each solvent at methyl tertiary butyl ether (MtBE) are not shown because water solubility
is less than the 10 wt% water assumed for the feed.
100 ∘ C. Areas calculated with the single-pass PV prediction spread-
sheet are presented in Fig. 12, normalized by the area required
when ethanol is the solvent and pP = 0 (Amin = 23.9 m2 ). The mem- 12.1 kPa, markedly higher than 50% increase caused by pP = 2.7 kPa
brane was assumed to deliver the same water permeance (2000 for ethanol. At the other end of the range, a 50% area increase
GPU) and selectivity (2000) for all solvents. The areas associated is predicted for methanol when pP is only 1.5 kPa. This trend is
with four permeate pressure conditions are shown as individual disrupted only for DMF as the area to remove water from DMF rises
vertical bars for each solvent, pP = 0, 1, 2.7, and 5 kPa. The perme- above that of even methanol at higher permeate pressures. Similar
ate pressure of 2.7 kPa is the pressure that causes a 50% increase in trends in area were calculated for an equivalent recirculating batch
area for ethanol at this temperature and is included as a reference PV system, also operated at 100 ∘ C.
for the other solvents. The other two non-negligible pressures of A change in solvent will also impact the concentration of sol-
1 and 5 kPa were selected to provide a range of permeate condi- vent in the permeate. The permeate solvent concentrations corre-
tions for comparison. MIBK and MtBE are not included in Fig. 12 sponding to the areas and conditions used to generate Fig. 12 are
because 10 wt% water exceeds the solubility limit of each at 100 displayed in Fig. 13. Here the trends are not as obvious as for area
∘ C. It should be noted that 100 ∘ C is above the normal boiling point
because the average solvent permeate concentration depends pri-
of most of the solvents considered here, requiring operation of the marily on the product of the solvent partial pressure and mem-
PV system above atmospheric pressure to maintain the feed as a brane area. In this way, although DMF has the lowest partial pres-
liquid. If operating the system at atmospheric pressure, then tem- sure of all the solvents, 1-butanol is predicted to yield the lowest
perature would be limited to the minimum boiling point of the permeate concentration because the membrane area required for
mixture over a given concentration range. the water/DMF separation is about three times that required when
For negligible permeate pressure (left-most bar for each solvent), 1-butanol is the solvent. When pP = 0 (solid black bars in Fig. 13),
the membrane area required to reduce water from 10 to 0.5 wt% the permeate concentrations for all solvents are predicted to be in
decreases in the order: methanol > DMF > ethanol > 2-propanol > the vicinity of 1 wt%, ranging from 0.21 wt% 1-butanol to 2.9 wt%
acetone > acetonitrile > 1-butanol > THF. This order is the reverse methanol. The effect of operating with non-negligible permeate
of the order of decreasing water partial pressures displayed in pressures on permeate solvent follows the trend for membrane
Fig. 7. Thus, only when the solvent was DMF or methanol did the area with methanol predicted to reach 16 wt% in the permeate if
calculations predict more area was required than when ethanol pP = 5 kPa, further illustrating the impact of pP on many aspects of
was the solvent. Although not shown, the area calculated from Eqn system performance.
(18) using the geometric mean feed-side water partial pressure
provided a reasonable estimate of the spreadsheet calculations for
membrane area of each solvent at pP = 0 (the average difference Effect of solvent on membrane performance
between the area from Eqn (18) and the area from the spreadsheet The earlier-mentioned predictions for the effect of solvent on
calculation was 4.4%). Solvent/water systems requiring the least PV system performance assumed water permeance and selectiv-
amount of membrane area when pP = 0 showed the lowest relative ity were the same for all solvents. These assumptions are useful
increase in area due to increasing pP . THF is predicted to require to scrutinize the impact of solvent vapor pressure properties on
the least membrane area for this separation. A 50% increase in system parameters, but it is more probable that the solvent will
509

area when THF is the solvent is predicted to occur when pP rises to impact membrane performance, particularly selectivity. The effect

J Chem Technol Biotechnol 2020; 95: 495–512 Published 2019. This article is a U.S. Government work wileyonlinelibrary.com/jctb
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org LM Vane

solvents and membrane–solvent interactions, solvent permeance


will depend on the solvent and, therefore, so will 𝛼. For example,
Tang et al. and Huang et al. reported the strong effect of molecular
size on permeability through amorphous perfluoro polymers.45–47
In one study, the permeability of water in the amorphous perflu-
oro polymer Hyflon AD 60 was about ten times that of methanol,
100 times that of 2-propanol, and 200 times that of 1-butanol.45
Although zeolite membranes operate, in part, on a size sieving
mechanism and the NaA, T-type, and CHA zeolites possess pore
sizes that should reject most solvents, the permeance of organic
compounds with actual zeolite membranes does not always track
with molecular size. The presence of non-zeolitic defects and
non-selective grain boundary defects may result in transport
behavior that does not match that of zeolite pores.48 Sommer and
Melin tested NaA and T-type zeolite membranes, as well as amor-
phous silica membranes, for the dehydration of a variety of organic
solvents.49,50 For the solvents in Fig. 12, water/solvent permselec-
tivities for a NaA zeolite membrane varied from 400 for acetoni-
Figure 13. Effect of solvent in a binary water/solvent mixture on the con- trile up to 31 000 for methanol.50 Each of the four alcohols stud-
centration of solvent in the permeate stream of a single-pass pervaporation ied (methanol, ethanol, 2-propanol, and 1-butanol) was reported
(PV) operating at 100 ∘ C as a function of total permeate pressure as calcu-
to have a water/solvent selectivity of over 10 000, as did THF. The
lated with single-pass PV spreadsheet. Same conditions as described for
Fig. 12. selectivity fell to 2700 for acetone and fell further to 520 and
400 for DMF and acetonitrile, respectively. Based on those values,
the permeate concentrations for the alcohols and THF with that
NaA membrane would be better represented by the right-most
(red) bars in Fig. 14, whereas acetone would be the black bar, and
the permeate for DMF or acetonitrile would be closer to the 𝛼
= 200 gray bars. Thus, although DMF would have nearly the lowest
permeate concentration of the solvents if selectivity was indepen-
dent of solvent type, it is likely to produce one of the highest for
the selectivities reported by Sommer and Melin for a NaA zeolite
membrane. While there was a 75-fold range in selectivity for the
NaA membrane, the ratio of highest to lowest water permeance
for the eight solvents in Fig. 12 was only 3.3.50 Thus, the solvent
that elicits the lowest water permeance would require 3.3 times
as much membrane area as that with the highest water perme-
ance, all other factors being equal. The magnitude of this range
in required area is on the order of the ratios of membrane areas
between solvents simply due to the solvent properties as shown
in Fig. 12.
For membrane materials that sorb water and swell as water activ-
Figure 14. Effect of water/solvent selectivity (𝛼) on the concentration
of solvent in the permeate stream of a single-pass pervaporation (PV) ity increases, a change in solvent could impact performance of
operating at 100 ∘ C. Values calculated with single-pass PV spreadsheet with water-swelling materials by impacting the water activity experi-
the same conditions as described for Fig. 12 except variable selectivity and enced by the membrane, thereby affecting both (Pi /𝓁) and 𝛼. For
pP =0. example, Yave recently reported on the effect of feed water con-
centration on two commercial poly(vinyl alcohol) membranes fab-
of selectivity on permeate solvent concentration for selectivities of ricated with different degrees of crosslinking.19 A higher degree of
20, 200, 2000, and 20 000 is shown in Fig. 14 for the same separa- poly(vinyl alcohol) crosslinking resulted in a lower dependence of
tion conditions as applied in the previous figure, including a fixed (Pw /𝓁) and 𝛼 values on water content and higher selectivity and
water permeance of 2000 GPU, and with pP = 0. The solid black lower permeances than that of the less crosslinked poly(vinyl alco-
bars designate 𝛼 = 2000 and are the same as the black bars in hol). If the effect of water content on membrane performance is
Fig. 13. As indicated in Fig. 14 for each solvent, a 10-fold decrease known, then that could be factored into the estimates, possibly as
in 𝛼 results in approximately a 10-fold increase in permeate con- a geometric mean permeance in the rough calculations or as vari-
centration – until a 10-fold increase in concentration is not physi- able permeances in the spreadsheet calculation.20–22 Similarly, if
cally possible. For the lowest selectivity of 20, each solvent would the effect of temperature on permeances was parameterized, this
be present at over 10 wt% in the permeate and most of the sol- relationship could be added to the estimation spreadsheets.
vents would be nearly 10 wt% in the permeate even for 𝛼 = 200.
The average permeate concentration falls to about 0.1 wt%, about Designing a PV system for drying multiple water/solvent
1000 ppm solvent in water, when selectivity is 20 000. mixtures
For selective layer materials that minimally swell in water and If a facility generates multiple water/solvent mixtures that can be
in the solvents, (Pw /𝓁) should be relatively constant between sol- accumulated in batches, then a single recirculating batch PV sys-
510

vents. However, due to the differences in molecular size of the tem, or a single-pass PV system that processes separate batches

wileyonlinelibrary.com/jctb Published 2019. This article is a U.S. Government work J Chem Technol Biotechnol 2020; 95: 495–512
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Review of pervaporation and vapor permeation process factors www.soci.org

of water-contaminated solvent, could be contemplated to process this work is that the process outcomes depend on the full range of
the mixtures.51 Batch systems provide more flexibility since time process parameters under non-ideal conditions.
is an additional design degree freedom.52 The processing limita-
tions and specifications of the range of mixtures would need to be
considered in designing the PV unit. Before engaging in a detailed ACKNOWLEDGEMENTS
process design, the rough design (membrane area requirements This work was conducted under the US Environmental Protection
and cycle times) for each mixture could be estimated with the Agency’s (USEPA’s) Sustainable & Healthy Communities National
rough estimation equation and evaluated more thoroughly with a Research Program. The views expressed in this article are those of
spreadsheet estimator or process simulator. The maximum perme- the author and do not necessarily represent the views or policies of
ate pressures could then be predicted along with the condenser the USEPA. Any mention of trade names, products, or services does
temperatures based on bubble points of the permeate composi- not imply an endorsement by the author, the US Government,
tions at the estimated pressures. or the USEPA. The USEPA and its employees do not endorse any
commercial products, services, or enterprises.

CONCLUSIONS Supporting Information


The performance of PV and V·P membranes is often boiled down Supporting information may be found in the online version of this
to flux and separation factor for a given set of process conditions. article.
In some cases, a single parameter has been used to compare mem-
branes. Unfortunately, test and literature data reported for mem-
REFERENCES
branes may not be directly useful for membrane system end-users 1 Cseri L, Razali M, Pogany P and Szekely G, Chapter 3.15 - organic sol-
because such systems rarely operate at the water concentration, vents in sustainable synthesis and engineering, in Green Chemistry,
temperature, or permeate pressure reported. In addition, test data ed. by Török B and Dransfield T. Elsevier, Amsterdam, pp. 513–553
may not be available for the solvent being considered for dry- (2018).
2 Overview of the 2015 Definition of Solid Waste Final Rule. Retrieved
ing. While coupon/module testing is recommended for identifying
25 October 2019, from https://19january2017snapshot.epa.gov/
complications and preparing a final system design, rough calcula- sites/production/files/2015-08/documents/dsw_fnl_rul_brfng_
tions can be performed with limited data to arrive at an estimate 012215.pdf (2015).
of membrane area requirements and process performance. 3 USEPA, Selection of Industry Sectors, Chemicals and Func-
Even for the single binary system of water/ethanol, when tions in the Remanufacturing Exclusion: Background Doc-
ument in Support of the Definition of Solid Waste Rule,
operated under a non-negligible permeate pressure, a change Document # EPA-HQ-RCRA-2010–0742-0012, https://www.
in one process condition resulted in changes in all process regulations.gov/document?D=EPA-HQ-RCRA-2010-0742-0012
outcomes including: membrane area, permeate concentration, (2011).
solvent recovery, permeate condenser temperatures, or heating 4 USEPA, Benefits of the Remanufacturing Exclusion: Background Doc-
ument in Support of the Definition of Solid Waste Rule, Docu-
requirements. In other words, when describing PV or V·P process
ment # EPA-HQ-RCRA-2010–0742-0011, https://www.regulations.
performance: ‘everything is connected to everything else’ – an gov/document?D=EPA-HQ-RCRA-2010-0742-0011 (2011).
expression of the Haida people to describe the interdependency 5 Amelio A, Figueroa Paredes DA, Degrève J, Luis P, Van der Bruggen B
and interconnectivity in the natural world. While the interdepen- and Espinosa J, Conceptual model-based design and environmental
dency in a PV/V·P process pales in comparison to that of nature, evaluation of waste solvent technologies: application to the sepa-
ration of the mixture acetone-water. Sep Sci Technol 53:1791–1810
the response of process performance to the parameters studied (2018).
herein indicates that PV/V·P processes are more complex than 6 Slater CS, Savelski M, Hounsell G, Pilipauskas D and Urbanski F, Green
generally portrayed. design alternatives for isopropanol recovery in the celecoxib pro-
The simplified equations developed herein proved useful in cess. Clean Techn Environ Policy 14:687–698 (2012).
7 Savelski MJ, Slater CS, Tozzi PV and Wisniewski CM, On the simulation,
estimating the minimum membrane area requirements when economic analysis, and life cycle assessment of batch-mode organic
the system is operated under negligible permeate pressure. The solvent recovery alternatives for the pharmaceutical industry. Clean
spreadsheet calculations enabled predictions for non-negligible Techn Environ Policy 19:2467–2477 (2017).
permeate pressures in a cross-flow configuration. The permeate 8 Vane LM, Separation technologies for the recovery and dehydra-
pressure threshold for raising required membrane area by 50% tion of alcohols from fermentation broths. Biofuels Bioprod Biorefin
2:553–588 (2008).
was lowest for the lowest feed temperature, lowest retentate water 9 Vane LM, Review: membrane materials for the removal of water from
concentration, and highest selectivity. A method of estimating the industrial solvents by pervaporation and vapor permeation. J Chem
maximum practical permeate pressure was developed with values Technol Biotechnol 94:343–365 (2019).
that mirror those identified with the spreadsheet based on the 50% 10 Wynn NP, Pervaporation: vapor permeation, in Encyclopedia of Chem-
ical Processing, ed. by Lee S. Taylor & Francis Group, New York, pp.
area increase threshold. 2031–2051 (2006).
While each of the membrane materials is expected to be capa- 11 Underwood AJV, Fractional distillation of multicomponent mixtures.
ble of removing water from almost all the targeted solvents, the Chem Eng Prog 44:603–614 (1948).
specific solvent was shown to greatly affect process outcomes. As 12 Henley EJ and Seader JD, Equilibrium-stage Separation Operations in
a result, although a single PV or V·P system could be contemplated Chemical Engineering. John Wiley & Sons, New York (1981).
13 Stichlmair JG and Fair JR, Distillation: Principles and Practices. Wiley-VCH,
to dry batches of different solvent/water mixtures, process param- New York (1998).
eters would need to be tailored to each batch to account for the 14 Skiborowski M, Recker S and Marquardt W, Shortcut-based optimiza-
effect of the solvent and any other processing difference, such as tion of distillation-based processes by a novel reformulation of the
initial water content and required final water content for reuse, feed angle method. Chem Eng Res Des 132:135–148 (2018).
15 Doherty MF and Malone MF, Conceptual Design of Distillation Systems.
on the process. In addition to being more flexible, recirculating McGraw-Hill, Boston, MA (2001).
batch PV processes were predicted to be less sensitive to perme- 16 Baker RW, Membrane Technology and Applications, 3rd edn. John Wiley
511

ate pressure increases. However, the overarching observation from & Sons, Chichester (2012).

J Chem Technol Biotechnol 2020; 95: 495–512 Published 2019. This article is a U.S. Government work wileyonlinelibrary.com/jctb
and is in the public domain in the USA.
10974660, 2020, 3, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jctb.6264 by Pcp/Nicholas Copernicus , Wiley Online Library on [01/07/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org LM Vane

17 Vane LM and Alvarez FR, Hybrid vapor stripping-vapor permeation 37 Nasirzadeh K, Neueder R and Kunz W, Vapor pressures, osmotic and
process for recovery and dehydration of 1-butanol and ace- activity coefficients of electrolytes in protic solvents at different
tone/butanol/ethanol from dilute aqueous solutions. Part 1. Process temperatures. 2. Lithium bromide in ethanol. J Solution Chem
simulations. J Chem Technol Biotechnol 88:1436–1447 (2013). 33:1429–1446 (2004).
18 Vane LM and Alvarez FR, Effect of membrane and process characteris- 38 Herold KE, Radermacher R and Klein SA, Absorption Chillers and Heat
tics on cost and energy usage for separating alcohol-water mixtures Pumps. CRC Press, Boca Raton, FL (1996).
using hybrid vapor stripping-vapor permeation process. J Chem 39 Vane LM, Mairal AP, Ng A, Alvarez FR and Baker RW, Membrane
Technol Biotechnol 90:1380–1390 (2015). Technology and Research, Inc. and The United States of America
19 Yave W, Separation performance of improved PERVAP membrane and as represented by the Environmental Protection Agency, Separa-
its dependence on operating conditions. J Membr Sci Res 5:216–221 tion process using pervaporation and dephlegmation. US Patent
(2019). 6,755,975 (2004).
20 Buchaly C, Kreis P and Górak A, Hybrid separation 40 Vane LM, Alvarez FR, Mairal AP and Baker RW, Separation of vapor-
processes – combination of reactive distillation with membrane phase alcohol/water mixtures via fractional condensation using
separation. Chem Eng Process 46:790–799 (2007). a pilot-scale dephlegmator: enhancement of the pervaporation
21 Valentinyi N, Andre A, Haaz E, Fozer D, Toth AJ, Nagy T et al., Experimen- process separation factor. Ind Eng Chem Res 43:173–183 (2004).
tal investigation and modeling of the separation of ternary mixtures 41 ZEBREX: Zeolite Membrane Dehydration Technology. Retrieved 25
by hydrophilic pervaporation. Sep Sci Technol:1–17 (2019). October 2019, from https://www.m-chemical.co.jp/en/products/
22 Koester S, Roghmans F and Wessling M, Water vapor permeance: the departments/mcc/aquachem/product/1201038_8078.html#anc
interplay of feed and permeate activity. J Membr Sci 485:69–78 Docs (2019).
(2015). 42 Press Release: Mitsui Zosen Machinery & Service and Mitsubishi
23 EN 15376:2014. Automotive Fuels – Ethanol as a Blending Component Chemical to Form Tie-up to Produce and Sell Zeolite Mem-
for Petrol – Requirements and Test Methods. European Committee branes. Retrieved 25 October 2019, from http://www.mes.co.jp/
for Standarization (CEN), Brussels (2014). english/press/2016/20160714.html (2016).
24 D4806-19a, Standard Specification for Denatured Fuel Ethanol for Blend- 43 Sander U and Soukup P, Design and operation of a pervaporation plant
ing with Gasolines for Use as Automotive Spark-Ignition Engine Fuel. for ethanol dehydration. J Membr Sci 36:463–475 (1988).
ASTM International, West Conshohocken, PA (2019). 44 Bausa J and Marquardt W, Shortcut design methods for hybrid mem-
25 Smallwood I, Solvent Recovery Handbook. Blackwell Science Ltd, Oxford brane/distillation processes for the separation of nonideal multi-
(2002). component mixtures. Ind Eng Chem Res 39:1658–1672 (2000).
26 Brüschke HEA, State-of-art of pervaporation processes in the chemi- 45 Huang Y, Ly J, Nguyen D and Baker RW, Ethanol dehydration using
cal industry, in Membrane Technology in the Chemical Industry, ed. hydrophobic and hydrophilic polymer membranes. Ind Eng Chem
by Nunes SP and Peinemann KV. Wiley-VCH Verlag, Weinheim Res 49:12067–12073 (2010).
(2006). 46 Huang Y, Ly J, Aldajani T and Baker RW, Membrane Technology
27 Rapin JL, The Betheniville pervaporation unit: the first large-scale and Research, Inc., Liquid-phase and vapor-phase dehydration of
productive plant for the dehydration of ethanol, in Proceedings organic/water solutions. US Patent 8,002,874 (2011).
Third International Conference of Pervaporation Processes Chemical 47 Tang J, Sirkar KK and Majumdar S, Pervaporative dehydration of con-
Industries. Bakish Materials Corporation, Englewood, NJ, USA, pp. centrated aqueous solutions of selected polar organics by a perflu-
364–378 (1988). oropolymer membrane. Sep Purif Technol 175:122–129 (2017).
28 Zeng W, Li B, Li H, Li W, Jin H and Li Y, Mass produced NaA zeolite mem- 48 Choi J, Jeong H-K, Snyder MA, Stoeger JA, Masel RI and Tsapatsis M,
branes for pervaporative recycling of spent N-methyl-2-pyrrolidone Grain boundary defect elimination in a zeolite membrane by rapid
in the manufacturing process for lithium-ion battery. Sep Purif Tech- thermal processing. Science 325:590–593 (2009).
nol 228:115741 (2019). 49 Sommer S and Melin T, Influence of operation parameters on the
29 Morigami Y, Kondo M, Abe J, Kita H and Okamoto K, The first large-scale separation of mixtures by pervaporation and vapor permeation with
pervaporation plant using tubular-type module with zeolite NaA inorganic membranes. Part 1: dehydration of solvents. Chem Eng Sci
membrane. Sep Purif Technol 25:251–260 (2001). 60:4509–4523 (2005).
30 Pervatech BV, Datasheet Module PVM-120-37-P. Retrieved from https:// 50 Sommer S and Melin T, Performance evaluation of microporous inor-
pervaporation-membranes.com/wp-content/uploads/2014/04/ ganic membranes in the dehydration of industrial solvents. Chem
Datasheet-PVM-120-37-P-version-2-5-2016.pdf (2019). Eng Process 44:1138–1156 (2005).
31 Néel J, Pervaporation, in Membrane Separations Technology: Principles 51 Wijmans JG, Kaschemekat J and Baker RW, Pervaporation apparatus
and Applications, ed. by Noble RD and Stern SA. Elsevier Science, and process. Patent Application WO 1993012865A1 (1993).
Amsterdam, pp. 143–212 (1995). 52 Genduso G, Luis P and Van der Bruggen B, Overcoming any configura-
32 Fahmy A, Membrane Processes for the Dehydration of Organic Com- tion limitation: an alternative operation mode for pervaporation and
pounds. Ph.D. thesis. Universität Hannover, Hanover, Germany vapour permeation. J Chem Technol Biotechnol 91:948–957 (2016).
(2002). 53 Góral M and Wiśniewska-Gocłowska B, IUPAC-NIST solubility data
33 Fahmy A, Mewes D and Ohlrogge K, Absorption-assisted pervaporation series. 86. Ethers and ketones with water. Part 5. C6 ketones with
for solvent dehydration. Desalination 149:9–14 (2002). water. J Phys Chem Ref Data Monogr 37:1575–1609 (2008).
34 Weast RC, CRC Handbook of Chemistry and Physics. CRC Press, Boca 54 Góral M and Wiśniewska-Gocłowska B, IUPAC-NIST solubility data
Raton, FL (1980). series. 86. Ethers and ketones with water. Part 1. C2–C5 ethers with
35 Takaizumi K and Wakabayashi T, The freezing process in methanol-, water. J Phys Chem Ref Data Monogr 37:1119–1146 (2008).
ethanol-, and propanol-water systems as revealed by differential 55 Maczynski A, Shaw DG, Goral M and Wisniewska-Goclowska B,
scanning calorimetry. J Solution Chem 26:927–939 (1997). IUPAC-NIST solubility data series. 82. Alcohols with water – revised
36 Flick EW, Industrial Solvents Handbook, 5th edn. William Andrew Pub- and updated: part 1. C4 alcohols with water. J Phys Chem Ref Data
lishing/Noyes, Park Ridge, NJ (1998). Monogr 36:59–132 (2007).
512

wileyonlinelibrary.com/jctb Published 2019. This article is a U.S. Government work J Chem Technol Biotechnol 2020; 95: 495–512
and is in the public domain in the USA.

You might also like