Skip to main content
EPA Author Manuscripts logoLink to EPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Jan 1.
Published in final edited form as: J Chem Technol Biotechnol. 2019;94(2):343–365. doi: 10.1002/jctb.5839

Review: Membrane Materials for the Removal of Water from Industrial Solvents by Pervaporation and Vapor Permeation

Leland M Vane 1
PMCID: PMC6436640  NIHMSID: NIHMS1524180  PMID: 30930521

Abstract

Organic solvents are widely used in a variety of industrial sectors. Reclaiming and reusing the solvents may be the most economically and environmentally beneficial option for managing spent solvents. Purifying the solvents to meet reuse specifications can be challenging. For hydrophilic solvents, water must be removed prior to reuse, yet many hydrophilic solvents form hard-to-separate azeotropic mixtures with water. Such mixtures make separation processes energy intensive and cause economic challenges. The membrane processes pervaporation (PV) and vapor permeation (VP) can be less energy intensive than distillation-based processes and have proven to be very effective in removing water from azeotropic mixtures. In PV/VP, separation is based on the solution-diffusion interaction between the dense permselective layer of the membrane and the solvent/water mixture. This review provides a state-of-the-science analysis of materials used as the selective layer(s) of PV/VP membranes in removing water from organic solvents. A variety of membrane materials, such as polymeric, inorganic, mixed matrix, and hybrid, have been reported in the literature. A small subset of these are commercially available and highlighted here: poly(vinyl alcohol), polyimides, amorphous perfluoro polymers, NaA zeolites, chabazite zeolites, T-type zeolites, and hybrid silicas. The typical performance characteristics and operating limits of these membranes are discussed. Solvents targeted by the U.S. Environmental Protection Agency for reclamation are emphasized and ten common solvents are chosen for analysis: acetonitrile, 1-butanol, N,N-dimethyl formamide, ethanol, methanol, methyl isobutyl ketone, methyl tert-butyl ether, tetrahydrofuran, acetone, and 2-propanol.

Keywords: Solvent reclamation, membrane materials, dehydration, pervaporation, vapor permeation, membrane-based separation

Introduction

Organic solvents serve a variety of functions in the manufacture of chemicals and materials in the worldwide economy. Industries that use organic solvents provide economic opportunity in their communities, but the life cycles of industrial solvents – encompassing the production, transportation, use, and ultimate disposal/destruction of the solvent – introduce sources of emissions that potentially impact human health and the environment in those communities1, 2. According to the principles of Green Chemistry and Green Engineering, designing (or redesigning) products and processes to avoid the need for solvents is the preferred means of preventing pollution associated with solvent use35. If solvent use cannot be avoided in a process, the Green Chemistry/Engineering principles advocate employing the least amount of the safest solvent that delivers the requisite performance. In many situations, the life cycle impact of using even the safest solvents can be reduced by recovering and reusing the solvents2. For example, the Green Engineering Program of the United States Environmental Protection Agency (USEPA) identified solvents used in four industrial sectors as a potential opportunity to obtain large environmental benefits should those solvents be reclaimed and reused6. The solvents were used in high volumes and the production and disposal of these solvents were determined to require large amounts of energy7, 8. In 2015, the USEPA finalized a new Definition of Solid Waste Rule intended, in part, to encourage the reuse/re-processing of 18 higher-value hazardous spent industrial solvents used in four manufacturing sectors: pharmaceuticals, paints and coatings, plastics and resins, and basic organic chemicals9. The 18 solvents targeted for reclamation are listed in Table 1, grouped as non-polar, polar aprotic, and polar protic types of solvents. Although these 18 solvents are only a small fraction of the many organic solvents used in industry, they represent a cross-section of many types of organic solvents: aliphatic and aromatic hydrocarbons, chlorinated aromatic and aliphatic hydrocarbons, ethers, ketones, amides, nitriles, and alcohols.

Table 1.

Solvents covered by remanufacturing exclusion of the 2015 USEPA Definition of Solid Waste Rule.

Non-Polar Solvents Polar Aprotic Solvents Polar Protic Solvents
n-Hexane* Methyl tert-butyl ether (MtBE)* 1-Butanol*
Cyclohexane* Chlorobenzene* Methanol
1,2,4-Trimethylbenzene* Chloromethane Ethanol*
Toluene* Dichloromethane*
Ethylbenzene* Chloroform*
Xylenes* Methyl isobutyl ketone (MIBK)*
Tetrahydrofuran (THF)*
Acetonitrile*
N,N-Dimethylformamide (DMF)
*

Solvent forms an azeotrope with water (azeotrope information from Handbook of Organic Solvent Properties288 and Nishikawa et al.289)

Considered a Hazardous Air Pollutant (HAP) by the USEPA290

The technical challenge to solvent reuse/reprocessing is the application of suitable separation technologies to recover those solvents from mixtures with other processing materials and to purify them to quality specifications. Contaminants to be removed may include suspended solids, dissolved solids (salts, higher molecular weight soluble species), other solvents or oils, and water10. This review is focused on the reduction of water contamination in selected industrially-important solvents. Water may become part of the spent solvent in a variety of ways: added intentionally to alter the property of the solvent system, introduced in a washing step, produced via an intentional/unintentional reaction, or simply absorbed from the atmosphere.

Despite the wide range of solvent types represented in Table 1, most of these solvents, 15 of the 18, form difficult-to-separate azeotropic binary mixtures with water, as indicated by an asterisk next to the name of the solvent in the table. An azeotrope exists when a vapor phase in equilibrium with a liquid phase has the same composition as the liquid phase11. Under normal circumstances, the vapor phase will have a different composition than the liquid phase – the property that enables separation by evaporation and condensation. As a result, concentrating solvents beyond an azeotropic composition by simple evaporation/condensation, the basis for batch solvent stills and conventional distillation processes, becomes impossible. Fortunately, water is sparingly soluble in low polarity solvents, making it less likely that water will contaminate such solvents. Due to the partial miscibility found in many solvent/water systems, the liquid phases at azeotropic compositions often consist of two phases. In such cases, the azeotrope is referred to as “heterogeneous.” If the azeotropic liquid is a single phase, then it is referred to as a “homogeneous” azeotrope. In fact, 10 of the 18 targeted solvents solubilize less than 0.2 wt% water, as detailed in Table 2. In this group, the water solubility in the solvents ranges from only 0.0090 wt% water for n-hexane up to 0.18 wt% water for dichloromethane12. Also shown in Table 2 are the solubilities of these solvent in water.

Table 2.

Properties of solvents that exhibit sparing solubility in water and sparing solubility of water in the solvent, listed in general order of increasing polarity.

Solvent CAS No. MW¥
(g/mol)
Boiling Point¥
(°C)
Density @ 25°C¥
(g/mL)
Water in Azeo-trope @ 1 atm§
(wt%)
Temp. of Azeo-trope @ 1 atm§
(°C)
Solubility of water in solvent at 25°C
(wt%)
Solubility of solvent in water at 25°C
(wt%)
n-Hexane 110-54-3 86.18 68.7 0.656 6 62 0.0090 0.00095
Cyclohexane 110-82-7 84.16 80.7 0.773 9 70 0.016 0.0067
1,2,4-trimethylbenzene (pseudocumene) 95-63-6 120.19 168.9 0.859 23.4£ 97£ 0.038 0.0057
Toluene 108-88-3 92.14 110.6 0.865 20 85 0.054 0.057
Ethyl benzene 100-41-4 106.17 136.2 0.865 33 92 0.044 0.018
Xylenes (mixed) 1330-20-7 106.17 140 0.862 37 93 0.045 0.011
Chlorobenzene 108-90-7 112.56 131.7 1.101 28 90 0.033§ 0.054
Chloromethane 74-87-3 50.49 −24.2 N/A N/A 0.54**
Dichloromethane 75-09-2 84.93 39.8 1.318 1 38 0.18 1.3
Chloroform (trichloromethane) 67-66-3 119.4 61.2 1.480 3 56 0.087 0.78
£

From Nishikawa et al.289

¥

Yaws’ Critical Property Data for Chemical Engineers and Chemists, Knovel291, accessed via https://app.Knovel.com on 01/10/2018

§

Handbook of Organic Solvent Properties via Knovel288

Yaws’ Handbook of Properties for Aqueous Systems via Knovel12

**

At 1 atm partial pressure of solvent

The remaining 8 solvents from Table 1 are more hydrophilic than those in Table 2, dissolving 1 wt% or more of water, with five being fully miscible with water (see Table 3). Even in this subset of more hydrophilic solvents, 6 of the 8 form azeotropes with water. Thus, the solvents in Table 3 are both prone to accumulating water and form mixtures with water that are challenging to separate. As a result of these features, the Table 3 solvents were selected for further analysis with regard to the types of materials and processes that are suitable for removing water. Additionally, acetone and 2-propanol will be considered because they are representative of the types of solvents in Table 3 and data for their dehydration is often reported alongside those of the other solvents. Acetone, like MIBK, is a ketone, but, unlike MIBK, acetone does not form a water azeotrope and is fully miscible with water. 2-propanol adds a branched azeotrope-forming lower alcohol to the group of alcohols.

Table 3.

Physical properties of solvents targeted for reclamation that are miscible with water or exhibit a solubility of water in the solvent of 1 wt% or greater.

Solvent CAS No. MW¥
(g/mol)
Boiling Point¥
(°C)
Density @ 25°C¥
(g/mL)
Water in Azeo-trope @ 1 atm§
(wt%)
Temp. of Azeo-trope @ 1 atm§
(°C)
Solubility of water in solvent at 25°C
(wt%)
Solubility of solvent in water at 25°C
(wt%)
Acetonitrile 75-05-8 41.05 81.6 0.779 16 76.0 Miscible
1-Butanol 71-36-3 74.12 117.7 0.806 42.0 93.0 20.2 7.34
N,N-Dimethyl formamide (DMF) 68-12-2 73.10 153.0 0.945 No Azeotrope Miscible
Ethanol 64-17-5 46.07 78.3 0.787 4.0 78.0 Miscible
Methanol 67-56-1 32.042 64.7 0.787 No Azeotrope Miscible
Methyl Isobutyl Ketone (MIBK), 4-methyl-2-pentanone 108-10-1 100.16 117.7 0.796 24.0 88.0 2.15 1.85
Methyl tert-Butyl Ether (MtBE) 1634-04-4 88.15 55.2 0.735 3.0 52.0 1.0 3.25
Tetrahydrofuran (THF), Oxolane 109-99-9 72.11 64.9 0.880 5.3 64.0 Miscible
Commonly used solvents similar to those above
Acetone, Propan-2-one 67-64-1 58.08 56.3 0.786 No Azeotrope Miscible
2-Propanol (isopropanol, IPA) 67-63-0 60.10 82.3 0.783 12.6 80.0 Miscible
¥

Yaws’ Critical Property Data for Chemical Engineers and Chemists, Knovel291, accessed via https://app.Knovel.com on 01/10/2018

§

Handbook of Organic Solvent Properties via Knovel288

Heterogeneous azeotrope

Yaws’ Handbook of Properties for Aqueous Systems via Knovel12

In 2008, this author reviewed the types of separation technologies that could be used to recover alcohols from water and then dry those alcohols in the production of alcohol-based biofuels from dilute fermentation broths13. The categories of technologies applicable to drying alcohol biofuels described in that review also apply here to the drying of hydrophilic industrial solvents, namely:

  • Distillation

  • Adsorption

  • Liquid-liquid Extraction

  • Pervaporation

  • Vapor Permeation

Separation processes and separating materials take advantage of the chemical and structural differences between the compounds to be separated. As noted above, the vapor-liquid equilibrium (VLE) behavior is the primary chemical difference used in distillation-based processes to drive the separation utilizing single or successive evaporation/condensation step(s). Subcategories of distillation applicable to solvent drying include fractionation, azeotropic distillation, extractive distillation, and pressure swing distillation. Several of these involve modifying the conditions or solution components to alter the normal VLE behavior of the solvent/water mixture. The VLE behaviors for binary mixtures of three lower alcohols with water (methanol, ethanol, and 2-propanol) are shown in Figure 1. At low alcohol concentrations, the vapor phase is significantly enriched in the alcohol relative to the liquid phase for all three of the alcohols – as indicated by the VLE curves lying above the no-VLE-separation diagonal line. As alcohol concentration increases, the distance between the VLE curves and the diagonal line decreases and, in the cases of ethanol and 2-propanol, the VLE curves cross the diagonal line. The point of intersection is the azeotropic concentration. According to Table 3, at 25 °C, the ethanol/water and 2-propanol/water azeotropes occur at 4.0 and 12.6 wt% water, respectively. As indicated in Figure 1 and Table 3, the methanol/water system exhibits no azeotrope, although the VLE difference becomes small at high alcohol concentrations.

Figure 1.

Figure 1.

Vapor-liquid equilibrium of three alcohol/water mixtures of interest. Curves based on NRTL thermodynamic model at 101.3 kPa system pressure (see Table S1).

Two additional parameters that can be leveraged in separation processes are molecular size and solvent characteristics, the latter represented by solubility parameters. The kinetic diameters of water and several solvents are listed in Table 4. By this measure, water is the smallest of the molecules considered here, allowing it to pass through sieving structures more readily than the larger solvent molecules. The interaction of a solvent with another material is frequently characterized based on solubility parameters, particularly for prediction of sorption, swelling, and miscibility. Hansen solubility parameters provide a three-dimensional description of the interaction properties of a solvent: a dispersion contribution (δd), polar contribution (δp), and hydrogen bonding contribution (δh)14. The more dissimilar the properties of two solvents, the more likely it is that a third material could be introduced to preferentially interact with one of the solvents relative to another and affect a separation. This additional material, termed a mass separating agent, could be an adsorbent, liquid absorbent, or membrane.

Table 4.

Kinetic diameters of common solvents, including several from Table 344.

Solvent Kinetic Diameter (nm)
Water 0.296
Methanol 0.380
Ethanol 0.430
Acetic acid 0.436
Acetone 0.469
2-Propanol 0.470
Methyl acetate 0.478
Trichloromethane 0.483
THF 0.486
Pyridine 0.496
2-Butanol 0.504
Methyl ethyl ketone 0.504

The solubility parameters for water and the hydrophilic solvents listed in Table 3 are plotted in Figure 2 and the values are provided in Table S3. The solubility parameters for the more hydrophobic solvents listed in Table 2 are included at the bottom of Table S3 for comparison. As is clear from Figure 2, the largest difference between water and the hydrophilic solvents of Table 3 is observed in the hydrogen bonding contribution parameter. Methanol is the most similar to water in this parameter. Therefore, methanol/water mixtures are more difficult to separate utilizing molecular size and chemical differences than other organic solvents. The Hansen solubility parameter with the smallest range for the Table 3 solvents and water is the dispersion contribution parameter, suggesting this molecular aspect has the least separation leverage. Overall, mass separating agents that possess a high hydrogen bonding contribution and a high polar contribution are likely to preferentially interact with water relative to the organic solvents.

Figure 2.

Figure 2.

Hansen solubility parameters for solvents listed in Table 3.

For drying hydrophilic solvents, distillation and adsorption are the most established processes10, 16, 17. To separate azeotropic mixtures using distillation, a third solvent or compound can be added as an entrainer to alter the VLE behavior. In the separation of ethanol/water mixtures, benzene was once used as an entrainer in azeotropic distillation, but has since been replaced by less hazardous solvents such as cyclohexane10, 18, ethylene glycol19, 20, hyperbranched polymers21, and glycerol22. Adsorption with molecular sieves supplanted azeotropic distillation for drying ethanol in the U.S. ethanol biofuel industry decades ago23. In adsorption, the solvent/water mixture, either as a vapor or liquid, is contacted with a hydrophilic solid adsorbent with water selectively sorbing to the solid, leaving a water-laden adsorbent and a water-depleted solvent stream. The spent sorbent would then undergo regeneration, utilizing a swing in temperature or pressure, or both, possibly combined with the introduction of a sweep stream, to desorb the water from the sorbent. Thus, adsorption is a cyclic process. As a result, if continuous operation is desired, then multiple adsorbent beds/columns are required so that regeneration of one bed can occur while another bed is being loaded. The most recent advances in adsorbents have come in the areas of metal-organic framework (MOF) materials (including the subcategory of zeolitic imidazolate frameworks (ZIFs)), carbon molecular sieve (CMS) materials, and in the fabrication of molecular sieves into more efficient structures2428. For water adsorption from solvents, however, zeolite-based molecular sieve particles, particularly sodium A-type (“NaA” or “4A”) and potassium A-type (“KA” or “3A”) aluminosilicates, are still the primary materials used10, 17, 29, 30.

Other drying processes that may be suitable for industrial solvents are: reacting, salting out, coalescing, and fractional freezing10. For smaller scale drying applications, hydrate formation through contact with salts has been used10, 15. For example, solid calcium chloride removes water from hydrophobic ethers, esters, and hydrocarbons, but the salt is too soluble in alcohols, amines, and ketones to work effectively as an adsorbent10. Similarly, a significant challenge for liquid-liquid extraction is finding an extractant with a high selectivity toward water relative to the hydrophilic solvent.

Unlike most of the other drying processes, the membrane technologies of pervaporation and vapor permeation (PV and VP) operate on principles that leverage differences in both molecular size and solubility properties to separate miscible mixtures. This feature enables PV and VP to separate mixtures that might not be easily separated otherwise. The remainder of this paper will focus on the principles of PV/VP separations and the membrane materials that have proven effective at removing water from hydrophilic solvents.

Solvent/Water Separation by Pervaporation and Vapor Permeation

The emerging membrane technologies of PV and VP have been investigated for separating a variety of mixtures of volatile liquids, most often involving solvent/water mixtures. PV and VP operate under very different principles than filtration-type membrane processes (i.e. particle filtration, microfiltration, ultrafiltration). For filtration, the membrane is constructed with a porous top layer to allow fluid to flow under an applied pressure gradient through the membrane. However, the membrane in PV and VP is constructed with a dense top layer material such that molecules can only pass through that layer by a sorption-diffusion mechanism, often referred to as “solution-diffusion” transport. Molecules in the feed fluid sorb into the dense layer material, diffuse through that layer, and then desorb from the opposite side of the dense layer. As noted earlier, solubility parameters can be used to identify differences in solvent characteristics. Dense layer materials can be chosen based on the solubility parameters of the membrane material and the solvents to leverage these differences to yield a sorption selectivity3140.

By selecting a dense layer material that exhibits a sorption-diffusion selectivity towards one component of the feed fluid, PV and VP are able to selectively separate that component from the other component(s) in the feed fluid. The material that passes through the membrane, the “permeate” fluid, would then be enriched, relative to the feed fluid, in the preferentially transported component. Thus, while filtration membranes cannot separate simple miscible solvent/water solutions, PV and VP membranes can. The other difference is the driving force for transport through a PV or VP membrane is the chemical potential gradient across the membrane, as compared to the applied pressure gradient that controls fluid filtration. The chemical potential gradient is usually represented by the partial pressure gradient of a species. As a result, to cause a component of the feed fluid to pass through a PV/VP membrane, the partial pressure of the species in the feed fluid must be higher than the partial pressure of that species in the permeated downstream fluid. Thus, PV and VP operate under the same fundamental principles. The difference between them is the feed fluid in PV is a liquid while the feed fluid in VP is a vapor. The permeate in both cases is a vapor. A simplified illustration of a PV/VP separation unit is shown in Figure 3. The permeate vapor is typically recovered via cooling to yield a permeate condensate.

Figure 3.

Figure 3.

Schematic diagram of a pervaporation or vapor permeation membrane process. Molecules are depicted as circles.

The objective of this project is to analyze the state of the science of PV and VP in the field of water removal from the hydrophilic solvents identified in Table 3. For a more complete overview of PV/VP processes, a number of references are suggested reading including the works of Néel41, Baker42, Brüschke43, Bowen et al.44, and Vane45. Many operating and design parameters affect the performance of a PV/VP system. As noted above, the properties of the dense layer material of a PV/VP membrane are often the most consequential to system performance. As is the case with most material-based technologies, numerous membrane materials have been developed and evaluated for solvent drying, but only a few are commercially available. A search of the SCOPUS™ abstract and citation database using the words “pervaporation/pervapouration and dehydration and solvent” yields over 5000 documents in the scientific literature and over 1700 patents, with most activity since 1980. The emphasis of this paper is the state of selective materials development for PV/VP for the dehydration of hydrophilic solvents, providing a synopsis rather than an in-depth review of each type of material. Both commercial and research materials will be discussed although emphasis will be given to materials that are commercially available. The membrane materials will be further differentiated into organic, inorganic, and mixed matrix material categories.

Terminology and Fundamentals

In this section, to establish a common frame of reference for the discussion of membrane material performance, the terminology used in reporting the performance of PV/VP membranes will be outlined. In typical PV/VP experiments, the compositions of the feed and permeate fluids are measured, as are the amount of permeate collected per unit time, the permeate pressure, and the temperature of the feed fluid. For VP experiments, the total pressure of the feed vapor must be known in addition to the parameters measured for PV experiments. The total flux is calculated as the mass of permeate collected per unit time divided by the active membrane area.

The preferred method of reporting the separation performance of a PV/VP membrane material is in terms that have been normalized by the partial pressure driving force and the thickness of the dense layer (): the molar permeability of each species (Pi) and the molar permselectivity (α). In cases where is unknown or difficult to discern, the permeance (Πi), the ratio of permeability to thickness, may be reported instead of permeability. For commercial membranes, permeance is the most appropriate measure of performance since may not be reported and is not an experimental variable as in research studies. These parameters are calculated as:

Pi=wiV J lpiF- piV MWi and Πi=Pil  Equation 1

Where J is the total flux (kg/m2·s), wiV is the mass fraction of species i in the permeate vapor, MWi is the molecular weight of species i (kg/kmol), and where piF and piV are the partial pressures (kPa) of species i in the feed fluid and permeate vapor, respectively. SI units for permeance are kmol/m2·s·kPa and those for permeability are kmol·m/m2·s·kPa. Common units for permeance are gas permeation units (GPU), with 1 kmol/m2·s·kPa = 2.99×109 GPU and 1 GPU defined as 10−6 cm3(STP)/cm2·s·cmHg. Common units for permeability are Barrer with 1 kmol·m/m2·s·kPa = 2.99×1015 Barrer and 1 Barrer defined as 10−10 cm3(STP)·cm/cm2·s·cmHg.

For VP calculations, the feed and permeate partial pressures are simply calculated as the product of the mole fraction of a species in the feed or permeate (xi or yi) times the total observed pressure for that stream (i.e. piF=xipF and piV=yipV). For PV, the permeate partial pressure is calculated in the same manner as for the permeate in VP. However, the partial pressures for components of the PV feed liquid are calculated as the partial pressure of a given species in a hypothetical vapor phase that is in equilibrium with the feed liquid phase (i.e. piF=xiγipisat). Such a calculation requires either experimental VLE data or thermodynamic models to calculate the activity coefficient (γi) and the saturated vapor pressure (pisat).

Molar permselectivity (i.e. “selectivity”) is the ratio of the molar permeabilities or permeances of two species through the membrane:

α12=P1P2=Π1Π2 Equation 2

Although the driving force-normalized performance parameters defined in Equations 1 and 2 are preferred, PV and VP results are often reported in terms of individual fluxes and the separation factor (β), where the latter is calculated from mass fractions or mole fractions in the feed fluid and permeate as:

β12=(J1/J2)w1F/w2F= w1V/w2Vw1F/w2F=(y1/y2)x1/x2 Equation 3

(It should be noted the literature contains a variety of symbols and terms for what are defined here as permselectivity and separation factor and the definition and notation of “selectivity” in each article should be carefully assessed.) For VP experiments with negligible permeate pressure, the separation factor is equal to the permselectivity. However, for PV, the two terms differ at least by the relative VLE selectivity due to the need for the activity coefficient and saturated vapor pressure terms in Equation 1. In other words, the PV separation factor is determined by both the membrane selectivity and the VLE selectivity. As a result, it is possible to have a membrane that shows no molar permselectivity (i.e.α =1), but would still register an appreciable PV separation factor. Further, some materials are reported as being “selective” for one compound over another based on a PV separation factor greater than 1, but which are actually permselective for the other compound. The separation of ethanol from dilute aqueous solutions with silicone rubber membranes is a case in point with ethanol/water separation factors greater than 1 observed but molar permselectivity values less than 1 calculated after accounting for the favorable enrichment of ethanol in the vapor by VLE46, 47. Thus, on a molar permselectivity basis, silicone rubber is modestly water selective.

The reason for the water selectivity of silicone rubber goes to the properties that determine permeability. Permeability is defined as the product of solubility and diffusivity (i.e. Pi = Si × Di)48. Thus, the molar permselectivity is the product of the ratio of solubilities and the ratio of diffusivities as:

α12=S1S2×D1D2 Equation 4

Ideally, both the solubility and diffusivity ratios in Equation 4 favor one of the species. However, in some cases, including the separation of ethanol/water by silicone rubber, the diffusion selectivity and solubility selectivity are antagonistic. The size of the penetrant primarily dictates diffusion. Due to the small size of water relative to organic solvents, diffusion selectivity usually favors water. As a result, a membrane material would have to exhibit a high solubility selectivity toward the organic compound in order for the solvent/water selectivity to be greater than one. Thus, membranes have an uphill battle being more permselective for an organic solvent relative to water. However, the reverse objective, selectively removing water from an organic solvent, has the advantage of a favorable water/solvent diffusion selectivity. Pairing that advantage with a favorable water/solvent solubility selectivity makes for a truly water-selective membrane.

Mass Transfer

A molecule migrating from the bulk feed fluid to the bulk permeate vapor encounters several resistances to migration: a boundary layer of feed fluid on the upstream face of the membrane, sorption from the feed fluid into the dense membrane layer, diffusion through the dense layer, desorption out of the dense layer, diffusion/convection through the support layers of the membrane (if present), and a boundary layer of permeate vapor on the downstream face of the membrane4850. These “resistances-in-series”, along with the partial pressure driving forces, determine the observed throughput and separation performance of the membrane. Ideally, the permeance of the dense layer represents the dominant mass transfer resistance, since it is the most selective step in the series. However, as the thickness of the dense layer is reduced to decrease the resistance of this layer and, thereby, increase the flux through the membrane, the other resistances become more consequential. In laboratory testing with thick films, the behavior of the dense selective layer is more easily discerned than from testing of thin, supported films such as the typical commercial membrane. When the membrane is packaged into a module, bulk feed maldistribution and total pressure gradients in vapor streams complicate the assessment of the properties of the dense membrane layer. Considering all the parameters that can affect mass transfer in membrane testing, care should be taken when comparing PV/VP data sets from different sources evaluated under dissimilar conditions.

Membrane Materials for Solvent Drying by Pervaporation and Vapor Permeation

In this section, the types of materials that are currently used or have been evaluated as the selective dense membrane layer for the drying of hydrophilic solvents will be reviewed. Commercially available materials will be emphasized.

1. Organic Polymeric Materials

1.1. Poly(vinyl alcohol)

The benchmark polymeric membrane material for solvent drying is poly(vinyl alcohol) (PVA)42, 5173. The repeat unit of PVA is relatively simple compared to most performance polymers, consisting of two carbon atoms in the backbone of the polymer with one pendant hydroxyl group attached to every second carbon atom (see Figure 4A). PVA is produced through the hydrolysis of poly(vinyl acetate) and can be obtained with different degrees of hydrolysis, indicating the fraction of acetate groups that have been converted to alcohol groups74. Hydrophilicity increases with increasing degree of hydrolysis. PVA has a glass transition temperature of about 80 °C, but melts above 200 °C and will begin to decompose before melting74. PVA membranes are cast from aqueous solution and the films are usually crosslinked in some fashion to reduce swelling of the membranes in PV/VP feed streams with high water activities. Swelling usually results in an increase in permeance, which is favorable, but a decrease in selectivity, which can be unfavorable. Excessive swelling may cause irreversible changes in the membrane, such as delamination from the support layer or defect formation. Crosslinking of PVA can be accomplished via an added functional component or post-casting thermal, chemical, and/or radiation treatment62, 69, 7579. Common crosslinking agents include dianhydrides, dialdehydes, and multifunctional acids. Depending on the type of bonds formed, the crosslinks can introduce chemical stability issues. For example, the hemiacetal bonds formed when a dialdehyde, like the commonly studied glutaraldehyde, is the crosslinking agent can be hydrolytically labile80. Thus, operation under high water activities and at elevated temperatures can lead to a loss of crosslinks. Commercial PVA-based membranes have a sensitivity to certain impurities in the solvent/water mixture, including aldehydes, organic acids, amines, acetals/ketals, mineral acids, and peroxides81.

Figure 4.

Figure 4.

Chemical structures of several polymers used to prepare the dense layer of PV/VP dehydration membranes: A) Poly(vinyl alcohol); B) Poly(carboxy methyl cellulose); C) Polyimide (Matrimid™ 5218 shown); D) Poly(benzimidazole); E) Teflon™ AF.

PVA-based membranes were among the first to be commercialized for PV/VP applications and are still commercially available from DeltaMem AG (Switzerland) as part of their PERVAP™ product line of flat sheet membranes. Although the PERVAP™ membranes have been available for many years, DeltaMem AG was only recently formed from a management buy-out of the membrane technology business of Sulzer Chemtech. Sulzer had previously acquired the membrane product line that dates to GFT Ingenieurburo für Industrieanlagenbau57, 58, 82. CM Celfa Membrantrenntechnik AG (Switzerland), which is now a special products unit in The Folex Group83, is mentioned as manufacturing composite membranes with a crosslinked PVA dense layer84. Similarly, AzeoSep™−2002 membranes from PetroSep Membrane Research (Canada) are said to be PVA on a poly(acrylonitrile) (PAN) support85. The longevity of PVA in the PV/VP industry reflects the functional stability and performance of this polymer. While the nature of the polymers used by a company in the manufacture of PV/VP membranes might be openly disclosed or deduced from patent or scientific literature, it is common for the companies to alter the materials over time and/or hold this information as a trade secret. For example, the selective layer in the Azeo Sep™ membranes offered by Petro Sep Corporation (and KmX Membrane Technology) for solvent dehydration is not disclosed and may no longer be based on PVA8688.

1.2. Cellulose Polymers

Cellulose-based polymers have an even longer history in the PV literature than PVA, with nitrocellulose collodion containers being used in the 1917 article credited with first using the word “pervaporation”89, 90. Many membranes with a permselective layer prepared from cellulose, cellulose esters, or cellulose ethers have been reported for PV/VP applications58, 9198. Cellulose acetate membranes are seen as the benchmark material in the field of gas separation for the removal of carbon dioxide from carbon dioxide/methane mixtures42, 99. In some situations, cellulose acetate filtration membranes can be used as a physical support material for thin permselective layers of another polymer100. It is likely there are commercial cellulose-based membranes available despite the lack of any specific mention of such membranes in descriptions of membrane offerings. For example, Membrane Technology and Research, Inc. (USA) recently reported on the performance of a cellulose ester-based membrane custom manufactured for them93. The structure of one type of cellulose, carboxymethyl cellulose (CMC), is shown in Figure 4B.

1.3. Polyimides

Polyimides represent another class of polymeric membrane materials that were developed for solvent dehydration decades ago and are commercially available today. The hydrophilicity of polyimides is attributed to hydrogen bonding interaction between water molecules and imide functional groups101. Early patents on the subject were assigned to Ube Industries, Ltd. (Japan) and the company continues to produce polyimide membranes102111. More recently, Whitefox Technologies Limited (United Kingdom) has emerged to offer polyimide membranes for ethanol dehydration and solvent/chemical drying112114. Polyimide polymer chemistry was also the basis for SifTek™ VP membranes developed by Vaperma Inc. for ethanol dehydration115117. Despite successful field demonstrations for the drying of biofuel ethanol and expansion of their membrane production line, Vaperma did not survive the worldwide economic downturn that began in 2008 and the resulting pause in investment in what was expected to be a rapid rise in infrastructure to produce cellulosic ethanol. Several membrane-related patents originally assigned to Vaperma are now assigned to Parker Filtration BV (The Netherlands). Although some of the literature reports and patents from companies producing polyimide dehydration membranes mention application to liquid streams, the commercial applications have been limited to VP with superheated vapor feeds. This may be associated with the reported reaction of water with imide groups in polyimides and hydrolytic scission reactions at high water activities and temperatures101, 118.

The scientific literature contains a variety of reports of polyimide materials for PV/VP dehydration of solvents. Some of these have been experimental polyimide/polyetherimide materials, while others have been commercially available polymers, such as P84™ (Evonik, Austria) and Matrimid™ (Huntsman Chemical, USA)119125. The chemical structure of Matrimid™ 5218 is shown in Figure 4C. In some cases, the polyimide is used in the preparation of the porous support layer for a dense selective layer of another polymer126, 127. Thermally rearranged (TR) polymers have recently been reported for solvent dehydration128, 129. TR polymers are formed when ortho-functional polyimides are thermally treated at temperatures up to about 450 °C, causing a condensation reaction to yield a heterocyclic, aromatic structure that is both thermally stable and resistant to swelling130. Classes of TR polymers include poly(benzoxazole), poly(benzimidazole) (PBI) (see Figure 4D), and poly(benzothiazole)128131. Whitefox Technologies Limited has been developing PBI-based membranes for gas separation applications132. A 2017 blog post from Lux Research indicates that Whitefox Technologies also offers PBI membranes for the removal of water from solvents114.

1.4. Other hydrophilic polymers

Aside from PVA, cellulose, and polyimides, the scientific and patent literature discloses numerous polymers for PV/VP solvent drying applications. Chitosan- and alginate-based polymers have received significant attention over several decades of research, but do not appear to be commercially available at present133135. Other polyelectrolyte materials have been widely studied for PV/VP applications, including those formed using a layer-by-layer deposition method136142. Blends of polymers, interpenetrating networks, and copolymers provide opportunities to synergistically combine properties of individual polymers. Example blends and copolymers for dehydration applications include PVA/hydroxyethyl cellulose143, PVA/CMC144, PVA/poly(acrylic acid)60, 145149, PVA/sodium alginate150, and PVA/poly(allylamine)-hydrochloride151, 152.

1.5. Amorphous Perfluoropolymers

Most, if not all, of the permselective polymer materials mentioned above exhibit some sorption selectivity toward water relative to the targeted organic solvents. This sorption of water often results in the swelling of the polymer matrix and a change in the permeability and selectivity. A newer class of polymers, amorphous perfluoropolymers (APFPs), shows little, if any, change in permeability or selectivity due to contact with water, or even most solvents, because of minimal swelling in water153. Examples of APFPs that have been studied for solvent dehydration include Teflon™ AF polymers (e.g. AF2400 and AF1600, Chemours, USA), Cytop™ (Asahi Glass, Japan), and Hyflon™ AD polymers (e.g. AD60 and AD40, Solvay Specialty Chemicals, Italy). These are random copolymers of tetrafluoroethylene and bulky dioxolenic units such as the poly(2,2-bis(trifluoromethyl)-4,5-difluoro-1,3-dioxole-co-tetrafluoroethylene) used in the Teflon™ AF polymers (see Figure 4E). The dioxole groups disrupt the crystallinity normally found in poly(tetrafluoroethylene) (PTFE) to yield amorphous polymers possessing high fractional free volumes that enable permeation154. The glass transition temperatures of most APFPs are in the range of 110 to 240 °C, meaning they are in a glassy state at most PV/VP operating temperatures155. APFPs are insoluble in organic solvents apart from perfluorinated solvents, such as perfluoropolyethers. As with perfluorinated polymers like PTFE, APFPs are more chemically stable than most polymers. Some membrane research groups and companies, such as Compact Membrane Systems and Membrane Technology and Research, are developing APFPs tailored for permselective membrane applications156158.

APFP materials are described as being hydrophobic155, however, they are permeable to water due to the high fractional free volume they possess. The lack of interaction with both water and solvents yields APFP membranes with stable PV/VP performance characteristics over a wide range of solvent/water compositions and conditions159. The nature of the free volume elements in APFPs favors the transport of water molecules compared to the larger solvent molecules. In addition, for gas separation, the clustering of water molecules in the free volume elements has been proposed as a hindrance to transport of permanent gas molecules160. Huang et al. reported the permeances of water, methanol, ethanol, acetic acid, 2-propanol, and 1-butanol from binary solvent/water mixtures decreased dramatically with increasing molecular size of the compounds (reported as critical volume) in Hyflon™ AD6084, 159. For example, for composite membranes with a ~0.5 μm thick layer of Hyflon™ AD60, permeances of water, methanol, ethanol, 2-propanol, and 1-butanol measured during PV testing were on the order of 1100, 100, 16, 10, and 7 GPU, respectively, resulting in water/alcohol permselectivities on the order of 11, 70, 110, and 160 for the 1-, 2-, 3-, and 4-carbon alcohols, respectively. Smuleac et al. reported water/ethanol and water/2-propanol permselectivities for an APFP material (CMS-3, Compact Membrane Systems, USA) that was stable for water concentrations ranging from 10 wt% to over 90 wt%. The water/ethanol and water/2-propanol permselectivities for the CMS-3 material were 20 and 60, respectively157. Separately, Compact Membrane Systems reported water/ethanol permselectivities for PV and VP conditions of 25 and 18, respectively,158 and a water/THF PV permselectivity of 164161. Tang et al. reported on the ability of CMS-3 membranes to separate water from polar organic compounds, including acetone, 1-butanol, ethanol, DMF, N,N-dimethylacetamide (DMAc), and N,N-dimethylsulfoxide (DMSO)153, 162, 163. The permeability of water through the CMS-3 material was about 1000 Barrer (3.34×10−13 kmol·m/m2·s·kPa). Huang et al. studied three APFP materials and reported performance differences between the Cytop™, Hyflon™ AD60, and Teflon™ AF1600 APFP materials for water/ethanol separation with the following trend in permselectivity: Cytop™ > Hyflon™ AD60 > Teflon™ AF1600 (110 > 70 > 27)84, 159. The trend in water permeance was the opposite of permselectivity, demonstrating a trade-off between permeability and selectivity that is common in permselective membranes. The permeability of these APFP increases with increasing polymer glass transition temperature (Tg)155 with Teflon™ AF2400 exhibiting both the highest Tg and highest gas/vapor permeability of the commercially available Cytop/Hyflon/Teflon AF polymers.

1.6. Emerging organic materials

Carbon molecular sieves (CMSs) are a class of amorphous materials exhibiting molecular sieving characteristics and have been developed for membrane-based separations164. CMS materials are the microporous carbon skeleton remaining after pyrolysis of an organic polymer precursor material. Despite their origin as organic polymers and the retention of a carbon backbone, CMS materials have more in common, from a permselective transport perspective, with inorganic molecular sieves. CMS membranes have received significant research attention in the last several decades though there is limited information available on their use for PV/VP dehydration applications165168. Nevertheless, Media and Process Technology, Inc. (USA) markets a ceramic-supported CMS membrane for solvent drying and claims a water/solvent separation factor of up to about 2,000 – likely for a water/acetone mixture (water/methanol and water/ethanol mixtures are mentioned under the same heading)169.

Graphene and graphene oxide are additional carbon-based materials with molecular sieving capabilities that show promise as permselective membrane materials170172. These graphene materials are two-dimensional, atomically thin platelets that have the potential to create layers that are thinner than typical polymeric membranes. A few applications of graphene/graphene oxide membranes have involved the separation of water/solvent mixtures173181. For example, in a recent paper, a graphene oxide membrane was reported to deliver a water/2-propanol separation factor of over 800 and a flux of over 4 kg/m2·h in PV experiments with a feed stream of 5 wt% water at 70 °C173, indicating both good permselectivity and throughput.

Polymers that have received significant research attention lately, and show promise for solvent dehydration, are the TR polymers and polymers of intrinsic microporosity (PIMs). PIMs, such as PIM-1, possess a molecular structure that imparts microporosity enabling them to exhibit permselective behavior that is similar to that of conventional microporous materials, like zeolites, but with greater ease of preparation182. PIM-1 has been described as being ethanol-selective based on observed ethanol/water separation factors greater than 1183. However, as was previously described for silicone rubber membranes, the ethanol/water separation factors for PIM-1 membranes were less than that of natural ethanol/water VLE, indicating the PIM-1 membranes were modestly permselective for water under the conditions evaluated.

All three of these types of emerging materials (CMS, graphene, and PIM) have found more application as fillers in polymers than as standalone films. Such mixed matrix materials will be explored in a later section.

1.7. Layering of polymeric permselective materials

The majority of commercial polymeric membranes are layered composites with a dense permselective layer coated on a porous mechanical support structure that usually consists of a microporous material coated on top of a non-woven support. This allows the dense permselective layer to be very thin – on the order of 1 μm. In some cases, several thin layers of the same material are overcoated to build up the dense permselective layer. A number of reports describe the potential advantages of stacking layers of different polymeric permselective materials to form the dense permselective layer of the composite membrane93, 127, 159, 184187. For example, Membrane Technology and Research, Inc. described a multilayer membrane consisting of a layer of APFP on top of a cellulose ester layer93. As noted above, APFPs do not swell in water or in solvents while cellulose-based polymers are hydrophilic and will swell in water. Thus, the cellulose ester layer, by itself, would exhibit a wide swing in selectivity and permeance as a function of water activity in the feed liquid it contacted - with low permselectivity and high permeance when water activity was high. Huang et al. reported that the PV water/ethanol permselectivity of the bare cellulose ester membrane dropped two orders of magnitude as water content in the feed increased, from almost 1,000 under low water conditions (5 wt% water) to less than 10 when the feed contained 86 wt% water93. The water permeance increased by a factor of four over the same water concentration range, while ethanol permeance surged almost three orders of magnitude. Adding the APFP layer on top of the cellulose ester membrane caused a reduction in water activity in the cellulose ester layer due to the added mass transfer resistance between the feed fluid and the cellulose ester, resulting in an improved selectivity for water - at the expense of permeance. In fact, for much of the water concentration range, the APFP/cellulose ester layered composite delivered a permselectivity higher than that of either of the two permselective layers individually. The order of the layers is important, however. If the cellulose ester layer had been on top of the APFP layer (and, thus, in contact with the feed fluid), the performance of the composite under high water activities would be similar to that of the APFP layer alone due to high water activities and swelling throughout the cellulose ester layer. Hua et al. confirmed the synergistic behavior of multi-layered membranes by coating APFP Teflon™ AF2400 onto a hydrophilic poly(etherimide) hollow fiber and using them for the VP separation of mixtures of water with 2-propanol, ethanol, and 1-butanol187, 188.

The behavior of membranes composed of multiple permselective layers is determined by the complex interrelationship between the activities of the permeating molecules and the activity-dependent transport properties of the membrane materials. In addition to membranes prepared with multiple permselective layers, there are other multilayer applications where an additional layer is added to aid in mass transfer rather than as a specific permselective layer. In some cases, a protective or a high permeability “gathering” or “gutter” layer is added to improve properties of the composite such as overall lifetime, selectivity (by filling pinhole defects and reducing the impact of support porosity), or throughput (by allowing thinner permselective layers)42. Development of any multi-layer membrane is complicated by adhesion issues that naturally develop when materials of different thermal expansion and swelling properties are layered together. Such issues can lead to defect formation and delamination.

2. Inorganic materials

2.1. Zeolites – microporous crystalline inorganic materials

As noted earlier, zeolite-based molecular sieve adsorbent particles are used to remove water from solvents and bioethanol. The same types of zeolite crystals that are effective in adsorbent particles have been utilized to fabricate permselective membrane materials. Making a zeolite membrane is much more challenging than making an adsorbent particle because the membrane requires growth of a thin, defect-free zeolite crystal layer on a porous support. The mechanical support is usually made by depositing layers of ever finer particles of γ- and/or α-Alumina onto a base ceramic filter, although porous metal supports have also been used189, 190. Zeolites are classified as microporous TO2 crystals with uniform pore sizes, where T represents tetrahedral framework atoms such as Si, Al, B, Ge, Fe, and P44, 191. The specific repeating crystal lattice atoms determine the dimensions and interconnectivity of pores and cavities as well as the sorption characteristics. Because of the types of atoms and bonds involved, zeolite crystals used for membranes are chemically and thermally stable and do not swell in either water or most organic solvents, imbuing them with stable performance characteristics over a wide range of conditions.

2.1.1. Linde Type A (LTA) Zeolites

Linde Type A (LTA) zeolites, aluminosilicates with a 1:1 ratio of silicon to aluminum, are the most common type of zeolite used in PV/VP membranes. The lattice structure of an LTA zeolite is shown in Figure 5A. When a tetravalent silicon atom in SiO2 is replaced by trivalent aluminum atom, a positive charge deficiency is created192. An increase in the aluminum content results in an increase in hydrophilicity and decreases in both hydrothermal and acid resistances193. The charge introduced by aluminum atoms is balanced by introducing monovalent or divalent cations, such as sodium, potassium, calcium, and magnesium ions. These “counterions” affect the sorption and diffusion properties of the zeolite material. When sodium ions are the counter ions in an LTA lattice, the aluminosilicate is referred to as “NaA” or “4A” zeolite and the maximum pore size is 0.42 nm191. NaA zeolites are the benchmark inorganic membrane material for PV/VP dehydration applications194. If potassium ions are the counterions, then the resulting “KA” or “3A” LTA zeolite possesses a pore size of roughly 0.3 nm. As noted in Table 4, the kinetic diameter of water is 0.296 nm44. As a result, water molecules are able to pass through the pores of both KA and NaA zeolites. Of the solvents listed in Table 3, the solvent with the smallest kinetic diameter, at 0.38 nm, is methanol44, 195. Because both water and methanol are theoretically small enough to enter the pores of NaA zeolites, separation based on molecular sieving is limited. However, at 0.42 nm, ethanol is slightly, but meaningfully, larger than methanol and about the same size as the pores of a NaA zeolite membrane. All the other solvents in Table 3 are larger than ethanol, thus restricting their entry into the pores of both KA and NaA zeolites relative to water. There are a variety of estimates and measures of molecular dimensions, some of which yield larger sizes for the solvent molecules than the kinetic diameters referred to here196.

Figure 5.

Figure 5.

Structures of several inorganic materials used as the dense layer in PV/VP dehydration membranes: A) LTA zeolite lattice structure (e.g. NaA zeolite); B) CHA zeolite lattice structure; C) Microporous silica (left) and Hybrid Silica (right). (Figures A and B from Database of Zeolite Structures233. Figure C reproduced from Chang et al.300 with permission from The Royal Society of Chemistry.)

Since the molecular shape is more complicated than a single dimensional parameter and since both molecules and zeolite lattices are somewhat flexible, kinetic diameters and pore sizes should only be used as rough indicators of molecular sieving characteristics. Fortunately, molecular sieving is just one of the separation actions taking place with zeolite membranes. Sorption and diffusion within the zeolite pores also factor into the selectivity of zeolite membranes197. In addition, Bowen et al. described how each species in a multicomponent feed can impact the transport of other species through the constrained environment within a zeolite channel44. The net result can be water/solvent permselectivities that are surprisingly large compared to the size difference. For example, water/ethanol permselectivities for NaA zeolite membranes are consistently over 1,000 and commonly over 10,000 in both VP and PV modes194, 198200. Selectivities for solvents larger than ethanol are at least as high as those for ethanol201. Even for water/methanol, separation factors over 1,000 have been reported with NaA membranes198200.

While NaA zeolites do not undergo significant swelling due to water sorption, hydrothermal stability issues have been reported for NaA LTA zeolites in PV of water/ethanol mixtures at high water levels in the feed, specifically 50 or 80 wt% water at 70 °C and 80 wt% water at lower temperatures202. Exposure of aluminosilicates to steam can cause hydrolysis of Si–O–Al bonds in the zeolite framework resulting in dealumination and the formation of aluminum oxide or aluminum hydroxide species outside of the framework203. Better stability in VP mode relative to PV was reported, possibly due to lower water activity in the superheated vapor202. NaA zeolites can also undergo dealumination in strong acid environments (pH<6), compromising the cage structure and molecular sieving character194, 204. Damage to NaA membranes under alkaline conditions (pH>8) has also been reported, indicating the need to operate NaA membranes under neutral pH conditions205. Thermal stability issues have been described for zeolite membranes, not due to changes in the zeolite framework, but due to thermal expansion differences between the zeolite layer and the porous support layer206.

Compared to hydrophilic polymeric membranes, LTA membranes are considered to have a greater operating range with respect to water concentration and temperature as well as a longer average lifetime. However, early zeolite membranes suffered from low permeance (low water flux) due to thick zeolite layers, low water selectivity due to grain boundary/intercrystalline defects, and high cost to manufacture due to the many steps involved in the formation of a zeolite membrane. These issues have been resolved through the deposition of a uniform layer of zeolite seed crystals onto the porous support followed by hydrothermal secondary crystal growth treatment that yields a thinner defect-free zeolite layer194, 203, 207. Grain boundary defects that cannot be eliminated through seeding/secondary growth may be addressed via post-crystallization treatment(s)203, 208, 209. The majority of zeolite membranes are in single-channel tube form, although multi-channel tubes and monoliths, as well as hollow fibers, are emerging194, 210213.

Reports of zeolite PV membranes began to appear in the patent and scientific literature several years after polymeric PV membranes. One informal measure of the timing of the emergence of zeolite PV membranes relative to polymeric PV membranes can be found in the papers presented at the International Conference on Pervaporation Processes in the Chemical Industry, a conference held from 1986 through 1995. The first paper on zeolite membranes, specifically NaA membranes for solvent dehydration, was presented at the seventh of these conferences in 1995 jointly by Yamaguchi University and Mitsui Engineering & Shipbuilding Co., Ltd. (Japan)214. A U.S. patent for NaA membranes assigned to Mitsui was granted in the following year and Mitsui claims their first commercial installation of zeolite membrane was in 1998215, 216. Over the ensuing 20+ years, several companies have marketed NaA zeolite membranes. Based on internet searches, literature references, and other information sources, NaA membranes and related PV/VP systems are available from at least nine sources in 2018, indicating a competitive market for this type of membrane. A list of commercial suppliers of NaA membranes and systems known to the author along with contact/website information is included in Supporting Information.

Firm costs for PV/VP membranes and modules are difficult to obtain, but zeolite membranes are consistently described as being about ten times more expensive than polymeric membranes, per unit active area217. The largest portion of the cost of zeolite membranes is attributed to that of the ceramic support218. The costs for both polymeric and inorganic materials should decrease as fabrication methods become more efficient and scale of production increases. In a crude simplification, inorganic membranes offer a longer lifetime that partially offsets the higher per unit area cost. The thermal stability of inorganic materials enables them to be used at temperatures that might not be appropriate for most polymeric materials. Higher temperatures translate into higher partial pressure driving forces which, in turn, results in lower membrane areas required. Finally, for some PV/VP systems, the cost of the membrane system can be a minor component of the overall capital and operating costs, especially if compressors, vacuum pumps, large heat exchanger areas, and, if a hybrid process, distillation columns are involved219, 220. All factors considered, the higher cost per unit area of inorganic membranes may not translate into a significantly higher total system cost, if at all221.

2.1.2. T-type and Chabazite Zeolites

As noted above, NaA-type membranes can be degraded by dealumination when contacted with feed fluids exhibiting a high water activity or acidic condition. To extend the operating range of zeolite membranes for solvent dehydration, more stable zeolite frameworks than LTA have been developed. At least two of these, T-type (or “Zeolite T”) and chabazite (CHA), are commercially available. T-Type zeolites are an intergrowth structure of erionite and offretite crystals with a silicon to aluminum ratio of 3:1 to 4:1 vs. the 1:1 ratio of LTA zeolites222. The lower aluminum content is believed to make T-type zeolites less hydrophilic and more acid-stable than LTA zeolites. In this regard, T-type membranes were found to be stable in 50 wt% acetic acid, but experienced dealumination in 0.1 M HCl193. Ji et al. recently reported stable performance over 20 days of operation for a T-type membrane module removing water from an ethanol solution containing 9.3 wt% water and 6.9 wt% acetic acid (pH ~3) at 70 °C212. The framework of T-type zeolite yields pores that are slightly larger (effectively 0.36 nm × 0.51 nm) than those of NaA223, 224. As a result, T-type zeolites have demonstrated a lower selectivity for water than NaA, particularly for methanol and ethanol193, 198. T-type zeolite membranes are manufactured by Mitsui Engineering & Shipbuilding Co., Ltd. and were disclosed in the patent literature for PV/VP applications a few years after NaA membranes198, 225.

CHA membranes have emerged more recently in the PV/VP arena and are manufactured by both Mitsubishi Chemical Corporation (Japan)226229 and Hitachi Zosen Corporation (Japan)230232. Chabazite is a naturally occurring zeolite exhibiting a psuedohexagonal unit cell with ellipsoidal cages interconnected via 8-member ring windows in the form of rhombohedral crystals192, 233, 234. The maximum pore size of the 3-dimensional pore matrix for dehydrated CHA zeolites is 0.38 nm, falling between that of KA and NaA LTA zeolites44, 191, 234. The lattice framework is shown in Figure 5B. Like LTA zeolites, CHA can be ion exchanged to yield a variety of forms with slightly different pore sizes, including CHA-Na, CHA-K, and CHA-Ca233. CHA-type zeolites can be prepared with different silicon to aluminum ratios, with a ratio as high as 11:1 reported, and using various crystallization conditions213, 229, 235, 236. CHA membranes with higher Si:Al ratios are claimed to impart resistance to organic acids that cannot be treated with NaA membranes213, 229. A CHA-type zeolite membrane with a Si:Al ratio of about 3 was reported to have a water/ethanol separation factor of over 270,000 for an 8 wt% water feed liquid at 77 °C237. The maximum calculable separation factor was constrained by the 50 ppm analytical detection limit for ethanol in the permeate. The water/2-propanol separation factor for the same membrane type was similarly high. The separation factor for water/methanol was not as high as that of ethanol or 2-propanol, but was still about 600 for feeds containing between 15 and 65 wt% water. For all three alcohols, the separation factor dropped at low water levels, particularly when less than 1 wt% water, falling to about 100 for methanol and 2-propanol and to about 5,000 for ethanol. The decline in water/alcohol separation factor was attributed to freer movement of the alcohols through zeolitic channels and non-zeolitic grain boundaries due to the relative absence of adsorbed water molecules that would block alcohol transport237.

As mentioned previously, a higher silicon to aluminum ratio is expected to impart improved hydrothermal and acid stability in zeolites. A high silica CHA-type zeolite (Si:Al of 7.5) was reported to achieve a water/ethanol PV separation factor of 500 for 86 wt% water at 70 °C229. Separation factors for larger solvents including 2-propanol, THF, acetone, and N-methyl-2-pyrrolidone (NMP) were higher than for ethanol, ranging from around 3,000 for water/THF to over 30,000 for water/2-propanol. As with all membranes, these separation factors vary with membrane fabrication methods and similar high silica CHA-type membranes were reported to have lower water/2-propanol separation factors of about 1,600 (molar permselectivity of 1,500) for 10 wt% water at 75 °C235. The water/methanol separation was not reported, but is expected to be of low selectivity because the high silica CHA was able to selectively permeate methanol from a mixture with acetone229. Thus, it appears the CHA-type membranes with higher Si:Al ratios have a lower water selectivity than CHA-type membranes with lower Si:Al ratios, an attribute that is most noticeable with the smaller alcohols.

2.2. Microporous amorphous inorganic materials

Microporous hydrophilic silica membranes, formed from tetraethyl orthosilicate (TEOS) using a sol-gel deposition method onto a porous alumina support, showed early promise for PV/VP solvent dehydration applications238241 but were eventually found to suffer from poor hydrothermal stability242244. Fortunately, researchers discovered that adding an organic linkage in the structure improved stability while still yielding a water-selective microporous material242, 245, 246. This type of material is referred to as “hybrid silica” or “organosilica”. The general structure of silica formed from TEOS and a hybrid silica prepared using bis(triethoxysilyl)ethane (BTESE) are shown in Figure 5C. The chemical composition and pore structure can be fine-tuned through alteration of sol components and cure conditions247252. Hybrid silica membranes are commercially available as the HybSi™ product line from Pervatech BV (The Netherlands)253, 254.

Hybrid silica materials have been reported to yield water/solvent separation factors that are generally lower than benchmark NaA zeolites - with water permeances/fluxes that are of the same order of magnitude than those of NaA membranes, under comparable conditions255. For example, BTESE-derived hybrid silica membranes have shown water/2-propanol PV separation factors between 210 and 4,400 for 10 wt% water at 75 °C255. Under similar feed conditions (10 wt% water at 70 °C), Moriyama et al. recently observed somewhat higher water/2-propanol separation factors of 1,000 and 7,000 for BTESE-derived membranes prepared with high and low sol acid molar ratios252. They also reported the performance of these BTESE-derived membranes for several other alcohol/water and solvent/water systems. The water/ethanol separation factors were 170 and 1,120 for the high and low acid molar ratios, respectively, with 10 wt% water at 50 °C. Water permeance for the lower selectivity membrane (~6,800 GPU) was several times greater than that of the higher selectivity membrane (~3,000 GPU). When hybrid silica membranes were prepared using a sol based on bis(triethoxysilyl)methane (BTESM), a water/ethanol separation factor of 890 was reported while that of water/1-butanol was 10,500245. The water/methanol separation factor was markedly lower, only 23. The water permeance of BTESE-based membranes was observed to increase with increasing water concentration and with decreasing temperature in tests with water/2-propanol mixtures252, 255. Kreiter et al. prepared hybrid silica membranes based on BTESE, BTESM, and a mixture of BTESE and methyltriethoxysilane, testing them using 5 wt% water in binary mixtures with four alcohols: methanol, ethanol, 1-propanol, and 1-butanol256. They reported the water permeance decreased with increasing alcohol chain length. However, the tests were not performed at the same temperature. In fact, the test temperature was increased as alcohol chain length increased, from 55 °C for methanol to 95 °C for 1-butanol. Based on the dependency of water permeance on temperature reported by Nagasawa et al., it is likely that much, if not all, of the change in water permeance with alcohol chain length observed by Kreiter et al. can be attributed to differences in the PV temperature255. This is borne out in the results of Moriyama et al. in that the water permeances observed for methanol and 1-butanol solutions, both tested at 50 °C, differed by only about 20%252.

Considering the known instability of microporous silica membranes, the developers of hybrid silica membranes carried out a wide range of long-term stability tests with the new hybrid silica membranes256258. For example, van Veen et al. subjected HybSi™ membranes to acidic conditions (concentrated acetic acid and dilute nitric acid), high temperatures (up to 190 °C), and strong aprotic solvents (e.g. NMP), some tests for as long as 1,000 days258. The main limitation reported was the need to maintain the pH between 2 and 8. Although the membranes were stable under acidic conditions, the organic linkages provide no improvement to the standard silica deterioration under alkaline conditions258. In the long-term tests, water/solvent selectivity of the membranes increased for a few weeks before stabilizing. Water flux for most membranes declined initially and took longer than selectivity to stabilize, requiring 1–12 months, when the flux had declined by about 50%. Pervatech BV notes limitations of use for their HybSi™ AR membranes as temperatures up to 150 °C and pH values between 0.5 and 8.5253.

2.3. Other Inorganic and Hybrid Materials

Researchers continue to develop and evaluate zeolite materials that offer the separation performance of NaA zeolites, but with greater stability under PV/VP conditions. Faujasite (FAU) type zeolites, claimed to have excellent durability, heat resistance, acid resistance, and water resistance, have been reported for solvent dehydration, although commercial availability is not apparent at this time259, 260. Hydroxy sodalite has also been proposed as a more stable alternative to NaA zeolites261, 262. As with polymers, zeolite blends and layering of zeolites is possible. For example, Mitsubishi Chemical Corporation was recently issued a patent for zeolite membranes fabricated from multiple zeolite crystals, either consisting of a mixture of zeolite crystals in one layer or two layers of different zeolites263.

As mentioned earlier, MOF materials are emerging as sorbents. These hybrid organic-inorganic structures are also being evaluated as permselective membranes, but reports of MOF-only membranes for solvent dehydration is limited. Most PV applications of MOF materials involve mixed matrix materials as described below.

3. Mixed Matrix Materials

Fabrication of high surface area, thin, defect-free membranes at a large-scale from molecular sieving materials such as zeolites, graphenes, CMSs, and MOFs can be quite challenging. One alternative is to disperse nanoscale or microscale particles of the sieving materials into a polymeric continuous phase to form a “mixed matrix membrane” (MMM)170, 264, 265. The objective in creating a MMM is to combine the desirable molecular sieving separation characteristics of the particulate dispersed phase with the desirable film-forming characteristics of the polymeric continuous phase. The final permeability properties are expected to fall between those of the two matrices, often mathematically described using empirical relationships42, 266, 267.

The number of MMM systems that have been studied for water removal from organic solvents is roughly equal to the number of organic polymeric materials (and polymer blends) studied for that separation multiplied by the number of particulate materials that are sorption-selective for water. Since both material categories are large, the number of applicable MMM systems is even larger. As might be expected, MMMs prepared from the two benchmark hydrophilic materials, PVA and NaA zeolite particles, have been studied for solvent dehydration by PV and VP265, 268272. As anticipated, the performance of a PVA/NaA zeolite MMM falls between that of 100% PVA and 100% NaA membranes. More recently, the emerging molecular sieving materials of CMS, MOF, and graphene have been used as the dispersed phase in water permselective MMMs273277.

Despite the promising performance characteristics of MMMs, they are not without limitations. As with the adage “a chain is only as strong as its weakest link,” the most prominent weakness of a specific MMM is generally that of either of the materials. For example, adding particles of an NaA zeolite to a continuous phase of PVA does not impart additional acid resistance to the NaA or additional water swelling resistance to the PVA. Despite the wide array of research publications on the topic of MMMs for water removal from organic solvents, it is not apparent that any water-selective MMM is commercially available for PV/VP applications - possibly due to the additive weaknesses in MMMs and the added complexity of MMM fabrication. Historically, the most commonly studied commercial MMM for PV, the Sulzer Chemtech PERVAP™ 1070 membrane, was actually designed to be solvent-selective for the recovery of organic compounds from water, containing hydrophobic ZSM-5 zeolite particles dispersed in silicone rubber278.

4. Other Materials in Membrane Modules

While the permselective membrane is the most important material in determining PV/VP separation performance, other materials in a membrane module must properly function under process extremes for the membrane to operate as intended. A critical material for all membrane formats is the porous support. Adhesion between the selective layer and the support must be maintained even when dimensional changes occur in either material due to swelling, thermal expansion, and pressure swings. At the same time, the support should present a low mass transfer resistance relative to that of the selective layer. As thinner selective layers are produced, the resistance of the support can become an important factor279. Membrane modules come in a variety of formats: hollow fiber, tubular, multichannel monolith/tube, plate & frame, and spiral wound. A set of materials are inherent to each type of modules and may include housings, adhesives (glues/tapes), sealants, gaskets/seals, turbulence enhancers, and feed or permeate flow spacers42. When considering whether an existing PV/VP module or system can be used under different operating conditions or even with a different solvent, the integrity of all materials of construction must be reviewed.

Overview of Performance and Operational Restrictions for Common PV/VP Membrane Materials

The commercially available materials and many of the emerging materials have proven to be effective in removing water from most, if not all, of the targeted solvents listed in Table 3. Each material will have limitations to the range of conditions under which they can be used. The known general limitations of common commercially available dehydration membrane materials are provided in Table 5. Seven types of dense layer materials are included in the table: PVA, polyimide, APFP, NaA zeolite, CHA zeolite, Zeolite T, and Hybrid Silica. A list of commercial suppliers of these materials/systems along with contact/website information is included in Supporting Information. Also provided in Table 5 are representative water permeance and water/ethanol permselectivity values based on PV performance reported for ethanol/water mixtures containing 5 to 15 wt% water at between 50 and 80 °C. The water/ethanol system was chosen because it is one of the most widely studied solvent/water systems in the PV/VP literature owing to the need for drying ethanol biofuels beyond the VLE azeotrope and because it is one of the second-most challenging permselective dehydration separation for the solvents in Table 3. As noted above, the most challenging permselective water/solvent separation is that of the water/methanol system. Water/methanol permselectivities for the seven materials listed in Table 5 are listed in the last column for comparison with those for water/ethanol. The water/solvent permselectivity for Table 3 solvents other than ethanol and methanol will be higher than that of ethanol. The water permeances for dilute water/ethanol solutions are reasonably representative of other water/solvent solutions. This is particularly true for those mixtures with water activities in the same region as those of 5–15 wt% water in ethanol or for materials that are insensitive to water activity, like the APFPs.

Table 5.

Performance and Limitations of Common PV/VP Dehydration Membrane Materials.

Material Reported PV Water Permeance¥ (Πw) (GPU) Water/Ethanol PV Selectivity¥ (α) Source of membrane for ethanol/water performance data General Limitations of Membrane Material Water/Methanol PV Selectivity§
Excluded Table 3 Solvents Max. Water Conc.,£ (wt%) Max. Temperature,£ (°C) Other
PVA 1,100294 4,800294
DeltaMem PERVAP™ 4101 tested by Clausthal Univ. of Technol.294 DMF£,81 30–50£ (depends on x-linking)81 100£,81 pH range: 5–8, Avoid (conditional): Aldehydes, mineral acids, sulfoxides peroxides, acetals, amides, amines81 30 (PERVAP 2201)295 to 250 (PERVAP 1510)296
Polyimide 1,100215 to 3,700297 57297 to 280215 Experimental, Yamaguchi Univ.297 and Mitsui Engg. & Shipbldg.215 None stated 70,102 None stated, at least to 150110, Vapor feed Polyimides generally limited to vapor feed for long-term applications110 ≈10x lower than ethanol (e.g. 6 to 30)¢
APFP 930158 to 2,900159 20158 to 110159 Compact Membrane Sys. CMS3158 and experimental Cytop and Teflon AF1600159 Possibly DMF (may depend on support) None stated, 100,158 None stated, at least 130,158 None stated 5158 to 1184
NaA 4,500198 10,000198 Experimental, Mitsui Engg. & Shipbldg. et al. None 20£,298 150£,298 pH range: 6–8, Avoid: Mineral & organic acids 2,100198
CHA 35,000237 120,000237 Experimental, National Inst. of Advanced Industrial Sci. and Technol. (AIST) None None stated, at least 50,235 None stated, at least 130,235 None stated 2,000237
Zeolite T 3,500198 2,500198 Experimental, Mitsui Engg. & Shipbldg. et al. None None stated, 100,193 None stated, at least 135,198 Avoid mineral acids at low pH (e.g. pH<2) 27198
Hybrid Silica 6,500299 230299 Experimental, Energy Research Centre of the Netherlands et al. None None Stated 150£,253 pH range: 0.5–8.5253 59299

Information in this table is a rough overview, losing some of the finer nuances of material compatibility and range of feed conditions, the reader is urged to contact a vendor or access the original literature reference to obtain more detailed information on a material for specific applications.

¥

Water permeance and water/ethanol permselectivity – as reported or calculated from flux data obtained, if possible, with commercial or pre-commercial membranes for ethanol/water mixtures containing 5–15 wt% water at between 50–80 °C. The values in the table were from tests covering the narrower operating range of 5 to 10 wt% water and 60 to 77 °C. 1 GPU = 3.34×10−10 kmol/m2·s·kPa

Based on the maximum tested in literature reports, not necessarily the limit of the material.

£

Based on a product description from the commercial supplier for long term use applications

§

Water/methanol permselectivity – as reported for methanol/water mixture containing 5–15 wt% water in the range of 40–65 °C

¢

Based on reported vapor permeation permselectivities for ethanol/water and methanol/water mixtures102.

All told, these seven commercially available membrane materials may be used to remove water from practically all the solvents of interest. Of the Table 3 solvents, only DMF has been reported as being problematic and only for two of the membranes listed in Table 5: PVA and APFP. This is most likely a limitation of the support materials employed in these composite membranes rather than of the PVA or APFP dense layers. For PVA membranes with a PAN microporous support layer, the PVA layer may be compatible with DMF, but the PAN support can be swelled by DMF, negatively impacting membrane integrity. The most significant limitation, particularly for hydrophilic polymers, is the maximum water content of the feed fluid. As indicated in Table 5, the reported limitation for water in the feed for PVA is 30–50 wt% with that limit dependent on the nature of the crosslinking in the PVA matrix. For polyimides, up to 70 wt% water has been demonstrated, but that was for a superheated vapor feed, liquid feeds are not advised. APFP membranes have the greatest water tolerance of the polymeric materials, with no reported limit on the water concentration in the feed. For zeolite materials with low Si:Al ratios, like NaA zeolites, high water activities and non-neutral pH values can cause hydrolysis and dealumination. As a result, NaA zeolites are limited to feeds containing less than 20 wt% water and to a pH range of 6–8. CHA and T-type zeolites have no stated limits on the feed water content, with reports of CHA membranes used at up to 50 wt% water and Zeolite T membranes at up to 100 wt% water. These two zeolites are also more acid resistant than the NaA materials. Hybrid Silica has no stated limit on the water content in the feed and is more acid tolerant than the zeolites. However, Hybrid Silica can break down under alkaline conditions, so 8.5 is reported as the maximum pH.

The stated long-term maximum operating temperature for commercial PVA membranes is 100 °C - all the other materials listed in Table 5 can be operated at higher temperatures. For commercial NaA zeolite and Hybrid Silica membranes, the stated long-term maximum operating temperature is 150 °C. For APFP, CHA zeolite, and Zeolite T, experimental results have been reported for temperatures as high as 130–135 °C, but the actual temperature limit has not been explicitly stated. Overall, PVA appears to be the most sensitive of the membrane materials, particularly to temperature and secondary compounds/contaminants. APFP, CHA, Zeolite T, and Hybrid Silica are the least sensitive to operating conditions and feed constituents. This feature not only defines the useful operating range for the membrane, but also factors into the risk of membrane damage due to process upsets or process shutdowns/startups.

As for separation performance, there is higher variability between the materials covered in Table 5 in terms of water/ethanol permselectivity than in terms of water permeance. The lowest water/ethanol selectivity of 20 was observed for a “CMS3” APFP material. Cytop™ APFP was reported to have a higher water/ethanol selectivity of 110. The water/ethanol selectivity of polyimides in PV mode is observed to be somewhat higher than that of the APFPs, in the 57 to 280 range. Hybrid silica is in this same range, yielding a water/ethanol selectivity of 230. Commercial PVA membranes deliver the highest water/ethanol selectivities of the polymeric membranes, well over 1,000 (α=4,800 for DeltaMem AG PERVAP™ 4101). As for the zeolites, Zeolite T is reported to have a water/ethanol selectivity in the same range as that of the PVA membranes while the selectivity increases to 10,000 for NaA membranes and up to 120,000 for CHA.

The water permeances for the membranes chosen as representative of the common membrane materials are in a tighter range than the respective selectivities, at least for the conditions selected in preparing Table 5. This is noteworthy given that permeance is roughly inversely proportional to the thickness of the dense layer and the dense layers of these assorted types of membranes can be quite different. The polymeric membranes are at the lower end of the range, with water permeances of between about 1,000 to 3,000 GPU. Zeolite T, NaA, and Hybrid Silica membranes delivered water permeances in the 3,500 to 6,500 GPU range. A CHA zeolite membrane exhibited water permeance of 35,000 GPU, far exceeding the other materials in Table 5. The amount of publicly disclosed data on CHA zeolite membranes with ethanol/water systems under the conditions outlined in Table 5 is limited. Additional testing is needed to determine if the high water/ethanol selectivity (α=120,000) and high water permeance (Πw=33,000 GPU) reported for experimental CHA membranes is routinely observed with commercial-grade CHA membranes. For example, results reported by Mitsubishi for a CHA membrane with Si:Al of 3.5, tested at a higher temperature of 100 °C with an ethanol/water mixture containing 15 wt% water, delivered a slightly reduced performance (α=75,000 and Πw=5,600 GPU) compared to the CHA membrane referenced in Table 5228. Since permeances are inversely related to the thickness of the dense layer, it is likely that permeances of future membranes of these same materials will be higher than those reported previously as manufacturers and researchers find ways to make thinner selective layers. Gains in permeance may come at the expense of lower selectivities as thinner dense layers become harder to make without defects and the mass transfer resistance of the support layer has a greater impact on overall mass transfer.

The ethanol/water separation performance characteristics of the seven membrane types listed in Table 5 are presented in Figure 6 in a pseudo-Robeson log-log plot of water/ethanol selectivity vs. water permeance. Each material is shown as a region in the graph. For PVA, NaA, T-type zeolite, Hybrid Silica, and CHA the regions are rectangles representing the selectivity and permeance values given in Table 5 for commercial or pre-commercial membranes with a ±50% span added to represent future improvements or losses in either selectivity or permeance. For APFP and Polyimide, the results from two versions of commercial or pre-commercial membrane materials were included in Table 5 and are shown in Figure 6 as extended rectangles. Additional details regarding the specific membranes, references, and performance data used to generate Figure 6 are provided in Table S2 of Supporting Information. The intention of Figure 6 is to qualitatively capture the typical performance differences between commercially available PV/VP membrane materials. Similar trends in performance have been reported through direct comparisons of two or more of these types of materials for water/solvent separations68, 280283.

Figure 6.

Figure 6.

Rough performance comparison of membranes prepared from the seven common PV/VP membrane materials listed in Table 5 removing water from ethanol/water mixtures by pervaporation for feeds containing 5–15 wt% water at temperatures of 50–80 °C. See text for explanation of the location and size of the region shown for each type of membrane.

As noted previously, methanol is more similar to water than is ethanol, in both size and chemical properties, making the selective removal of water from methanol more challenging than from ethanol. This is reflected in the significantly lower water/methanol permselectivities reported in Table 5 relative to those for water/ethanol, about 1–2 orders of magnitude lower. The CHA and NaA zeolite membranes captured in Table 5 were found to deliver the highest water/methanol selectivities of 2,000 to 2,100, but those are much smaller than the 120,000 and 10,000 selectivities reported for the water/ethanol separation. The lowest water/methanol selectivities, 5 to 11, were reported for APFP materials.

Status of, and Near-Term Prospects for, Dehydration Membrane Materials

In a 2013 PV/VP tutorial article, this author noted two large-scale PV/VP applications, alcohol biofuel drying and gasoline desulfurization, that held the near-term potential to enable these technologies to graduate from “emerging technology” status. While desulfurization by PV/VP is still in development, PV/VP technologies have gained a foothold as accepted options in the drying of bioethanol, either as a replacement for the molecular sieve system or to increase the capacity of an existing distillation-molecular sieve system. This acceptance is exemplified by several recent retrofits involving Whitefox Technologies Ltd. hollow fiber membranes systems in U.S. ethanol plants and in a recently announced marketing collaboration between ICM Inc., a developer of fuel ethanol technologies and a designer of many U.S. ethanol plants, and Mitsubishi Chemical Corporation to utilize Mitsubishi’s zeolite membranes in the fuel ethanol drying process227, 284287.

Overall, a robust and varied PV/VP membrane industry has developed. The membrane material with the greatest number of manufacturers and suppliers appears to be NaA zeolite (see Supporting Information). However, over the last decade or so, membranes made with several “newer” materials have become commercially available, including CHA and T-type zeolites, APFP, and hybrid silica. Membrane manufacturers are working to improve performance and stability while reducing the cost of their products. Thus, it is expected that the selectivity vs. permeance performance regions depicted in Figure 6 will shift over time to the right (i.e. higher throughput/flux) and up (higher selectivity). No doubt one or more of the emerging materials outlined in earlier sections will eventually become commercially available and warrant placement in Figure 6, joining PVA, polyimide, APFP, NaA zeolite, CHA zeolite, Zeolite T, and Hybrid Silica. Fortunately, as outlined in Table 5, the available PV/VP membrane materials can be used under a wide range of conditions to dry the hydrophilic solvents of interest outlined in Table 3. The design of batch and continuous solvent drying processes based on the PV/VP membrane materials detailed herein will be the subject of a future paper.

Supplementary Material

Supplement1

Acknowledgements/Disclaimer

This work was conducted under the USEPA’s Sustainable & Healthy Communities National Research Program. The views expressed in this article are those of the author and do not necessarily represent the views or policies of the USEPA. Any mention of trade names, products, or services does not imply an endorsement by the author, the United States Government, or the USEPA. The USEPA does not endorse any commercial products, services, or enterprises.

Abbreviations/Nomenclature

Abbreviations

APFP

amorphous perfluoropolymer

BTESE

bis(triethoxysilyl)ethane

BTESM

bis(triethoxysilyl)methane

CHA

chabazite

CMC

carboxy methyl cellulose

CMS

carbon molecular sieve

DMAc

N,N-dimethylacetamide

DMF

N,N-dimethylformamide

DMSO

N,N-dimethylsulfoxide

FAU

faujasite

GPU

gas permeation unit (10−6 cm3(STP)/cm2·s·cmHg = 3.344×10−10 kmol/m2·s·kPa)

IPA

isopropyl alcohol, 2-propanol

LTA

Linde Type A

MIBK

methyl isobutyl ketone

MMM

mixed matrix membrane

MOF

metal-organic framework

MtBE

methyl tertiary-butyl ether

NaA

sodium Linde Type A zeolite

NMP

N-methyl-2-pyrollidone

PAN

poly(acrylonitrile)

PBI

poly(benzimidazole)

PIM

polymer with intrinsic microporosity

PTFE

poly(tetrafluoroethylene)

PV

pervaporation

PVA

poly(vinyl alcohol)

STP

standard temperature and pressure

THF

tetrahydrofuran, oxolane

TR

thermally rearranged

USEPA

United States Environmental Protection Agency

VLE

vapor-liquid equilibrium

VP

vapor permeation

ZIF

zeolitic imidazolate framework

Roman symbols

Di

diffusivity of species i in membrane material, m2/s

J

total mass flux through membrane, kg/m2∙s

membrane thickness, m

MWi

molecular weight of species i, kg/kmol

pisat

saturated vapor pressure of species i at system temperature, kPa

piF

partial pressure of species i in the feed vapor or feed liquid, kPa

piV

partial pressure of species i in the permeate vapor, kPa

Pi

permeability of species i in membrane material, kmol∙m/m2∙s∙kPa

Si

solubility of species i in membrane material, kmol/m3∙kPa

wiF

mass fraction of species i in the feed vapor or feed liquid

wiV

mass fraction of species i in the permeate vapor

xi

mole fraction of component i in the feed vapor or feed liquid

yi

mole fraction of component i in the permeate vapor

Greek symbols

α

molar permselectivity (“selectivity”)

β

separation factor

γi

activity coefficient of species i

δ

overall Hansen solubility parameter, MPa1/2

δd

dispersion contribution to solubility parameter, MPa1/2

δh

hydrogen bonding contribution to solubility parameter, MPa1/2

δp

polar contribution to solubility parameter, MPa1/2

Πi

permeance of species i through membrane, kmol/m2∙s∙kPa

References

  • 1.Clarke CJ, Tu W-C, Levers O, Bröhl A, Hallett JP. Green and Sustainable Solvents in Chemical Processes. Chemical Reviews. 2018;118(2):747–800. [DOI] [PubMed] [Google Scholar]
  • 2.Cseri L, Razali M, Pogany P, Szekely G. Chapter 3.15 - Organic Solvents in Sustainable Synthesis and Engineering In: Török B, Dransfield T, editors. Green Chemistry: Elsevier; 2018. p. 513–53. [Google Scholar]
  • 3.Beach ES, Cui Z, Anastas PT. Green Chemistry: A design framework for sustainability. Energy & Environmental Science. 2009;2(10):1038–49. [Google Scholar]
  • 4.Anastas PT, Warner JC. Green Chemistry: Theory and Practice New York: Oxford University Press; 1998. [Google Scholar]
  • 5.Anastas PT, Zimmerman JB. Design through the 12 principles of green engineering. Environmental Science & Technology. 2003;37(5):94A–101A. [DOI] [PubMed] [Google Scholar]
  • 6.Overview of the 2015 Definition of Solid Waste Final Rule 2015. [Available from: https://www.epa.gov/sites/production/files/2015-08/documents/dsw_fnl_rul_brfng_012215.pdf].
  • 7.Selection of Industry Sectors, Chemicals and Functions in the Remanufacturing Exclusion: Background Document in Support of the Definition of Solid Waste Rule, Document # EPA-HQ-RCRA-2010–0742-0012, (2011).
  • 8.Benefits of the Remanufacturing Exclusion: Background Document in Support of the Definition of Solid Waste Rule, Document # EPA-HQ-RCRA-2010–0742-0011, (2011).
  • 9.Definition of Solid Waste; Final Rule, Docket EPA-HQ-RCRA-2010–0742; FRL-9728–5-OSWER, RIN 2050-AG62 (2015).
  • 10.Smallwood I Solvent Recovery Handbook. 2nd ed. Oxford: Blackwell Science Ltd; 2002. [Google Scholar]
  • 11.Henley EJ, Seader JD. Equilibrium-stage separation operations in chemical engineering. New York: John Wiley & Sons; 1981. [Google Scholar]
  • 12.Yaws CL. Yaws’ Handbook of Properties for Aqueous Systems. Accessed as online version available at: https://app.knovel.com/hotlink/toc/id:kpYHPAS006/yaws-handbook-properties/yaws-handbook-properties. Knovel; 2012. [Google Scholar]
  • 13.Vane LM. Separation technologies for the recovery and dehydration of alcohols from fermentation broths. Biofuels, Bioprod Bioref. 2008;2:553–88. [Google Scholar]
  • 14.Wypych G. Chapter 2. Fundamental Principles Governing Solvents Use Section 2.3 Basic Physical and Chemical Properties of Solvents. In: Wypych G, editor. Handbook of Solvents: ChemTec Publishing; 2001. [Google Scholar]
  • 15.Casey M, Leonard J, Lygo B, Procter G. Purification and Drying of Solvents Advanced Practical Organic Chemistry Boston, MA: Springer US; 1990. p. 28–42. [Google Scholar]
  • 16.Gavaskar AR, Olfenbuttel RF, Jones JA. Onsite Solvent Recovery, United States Environmental Protection Agency, Report No. EPA/600/R-94/026. Washington, DC: 1993. [Google Scholar]
  • 17.Bastidas PA, Gil ID, Rodrígueza G, editors. Comparison of the main ethanol dehydration technologies through process simulation 20th European Symposium on Computer Aided Process Engineering – ESCAPE20; 2010. [Google Scholar]
  • 18.Mahdi T, Ahmad A, Nasef MM, Ripin A. State-of-the-Art Technologies for Separation of Azeotropic Mixtures. Separation & Purification Reviews. 2015;44(4):308–30. [Google Scholar]
  • 19.Kiss AA. Distillation technology - still young and full of breakthrough opportunities. J Chem Technol Biotechnol. 2014;89(4):479–98. [Google Scholar]
  • 20.Kamihama N, Matsuda H, Kurihara K, Tochigi K, Oba S. Isobaric Vapor–Liquid Equilibria for Ethanol + Water + Ethylene Glycol and Its Constituent Three Binary Systems. Journal of Chemical & Engineering Data. 2012;57:339–44. [Google Scholar]
  • 21.Seiler M Hyperbranched polymers: New selective solvents for extractive distillation and solvent extraction. Sep Purif Technol. 2003;30:179–97. [Google Scholar]
  • 22.García-Herreros P, Gómez JM, Gil InD, Rodríguez G. Optimization of the Design and Operation of an Extractive Distillation System for the Production of Fuel Grade Ethanol Using Glycerol as Entrainer. Ind Eng Chem Res. 2011;50:3977–85. [Google Scholar]
  • 23.Swain RLB. Molecular sieve dehydrators: Why they became the industry standard and how they work In: Ingledew WM, Kelsall DR, Austin GD, Kluhspies C, editors. The Alcohol Textbook, 5th ed. Nottingham, UK: Nottingham University Press; 2009. [Google Scholar]
  • 24.Zhang K, Lively RP, Zhang C, Koros WJ, Chance RR. Investigating the Intrinsic Ethanol/Water Separation Capability of ZIF-8: An Adsorption and Diffusion Study. The Journal of Physical Chemistry C. 2013;117(14):7214–25. [Google Scholar]
  • 25.Shah M, McCarthy MC, Sachdeva S, Lee A, Jeong H-K. Current Status of Metal-Organic Framework Membranes for Gas Separations: Promises and Challenges. Ind Eng Chem Res. 2012;51:2179–99. [Google Scholar]
  • 26.Pimentel BR, Parulkar A, Zhou E-k, Brunelli NA, Lively RP. Zeolitic Imidazolate Frameworks: Next-Generation Materials for Energy-Efficient Gas Separations. ChemSusChem. 2014;7(12):3202–40. [DOI] [PubMed] [Google Scholar]
  • 27.Jiang J Recent development of in silico molecular modeling for gas and liquid separations in metal–organic frameworks. Current Opinion in Chemical Engineering. 2012;1:138–44. [Google Scholar]
  • 28.Lively RP, Chance RR, Kelley BT, Deckman HW, Drese JH, Jones CW, et al. Hollow fiber adsorbents for CO2 removal from flue gas. Ind Eng Chem Res. 2009;48(15):7314–24. [Google Scholar]
  • 29.Jović S, Laxminarayan Y, Keurentjes J, Schouten J, van der Schaaf J. Adsorptive Water Removal from Dichloromethane and Vapor-Phase Regeneration of a Molecular Sieve 3A Packed Bed. Ind Eng Chem Res. 2017;56(17):5042–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Hiltz J, Taylor Z, Baier M. Design of an Ethanol Dehydration System. University of Saskatchewan; 2008. [Google Scholar]
  • 31.Knozowska K, Kujawski W, Zatorska P, Kujawa J. Pervaporative Efficiency of Organic Solvents Separation Employing Hydrophilic and Hydrophobic Commercial Polymeric Membranes. J Membr Sci. 2018;564:444–55. [Google Scholar]
  • 32.Zou Y, Liu Y, Muhammed Y, Tong Z, Feng X. Experimental and modelling studies of pervaporative removal of odorous diacetyl and S-methylthiobutanoate from aqueous solutions using PEBA membrane. Sep Purif Technol. 2018;200:1–10. [Google Scholar]
  • 33.Li W, Luis P. Understanding coupling effects in pervaporation of multi-component mixtures. Sep Purif Technol. 2018;197:95–106. [Google Scholar]
  • 34.LaPack MA, Tou JC, McGuffin VL, Enke CG. The correlation of membrane permselectivity with Hildebrand solubility parameters. J Membr Sci. 1994;86(3):263–80. [Google Scholar]
  • 35.Yang Z, Zhang W, Li J, Chen J. Polyphosphazene membrane for desulfurization: Selectingpoly [bis (trifluoroethoxy) phosphazene] for pervaporative removal of thiophene. Sep Purif Technol. 2012;93:15–24. [Google Scholar]
  • 36.Mujiburohman M, Mahdi KA, Elkamel A. Predictive model of pervaporation performance based on physicochemical properties of permeant-membrane material and process conditions. J Membr Sci. 2011;381:1–9. [Google Scholar]
  • 37.Anim-Mensah Alexander R, Mark James E, Krantz William B. Use of Solubility Parameters for Predicting the Separation Characteristics of Poly(dimethylsiloxane) and Siloxane-Containing Membranes Science and Technology of Silicones and Silicone-Modified Materials. ACS Symposium Series. 964: American Chemical Society; 2007. p. 203–19. [Google Scholar]
  • 38.Lue SJ, Ou JS, Chen S-L, Hung W-S, Hu C-C, Jean YC, et al. Tailoring permeant sorption and diffusion properties with blended polyurethane/poly(dimethylsiloxane) (PU/PDMS) membranes. J Membr Sci. 2010;356:78–87. [Google Scholar]
  • 39.Ray SK, Sawant SB, Joshi JB, Pangarkar VG. Methanol Selective Membranes for Separation of Methanol-Ethylene Glycol Mixtures by Pervaporation. J Membr Sci. 1999;154(1):1–13. [Google Scholar]
  • 40.Nijhuis HH, Mulder MHV, Smolders CA. Selection of Elastomeric Membranes for the Removal of Volatile Organics from Water. J Appl Polym Sci. 1993;47:2227–43. [Google Scholar]
  • 41.Pervaporation Néel J.. In: Noble RD, Stern SA, editors. Membrane separations technology: Principles and applications Membrane Science and Technology Series. Amsterdam: Elsevier Science, B.V.; 1995. p. 143–212. [Google Scholar]
  • 42.Baker RW. Membrane technology and applications, 3rd ed. Sussex West, England: John Wiley & Sons, Ltd.; 2012. [Google Scholar]
  • 43.Brüschke HEA. State-of-Art of Pervaporation Processes in the Chemical Industry In: Nunes SP, Peinemann KV, editors. Membrane Technology in the Chemical Industry. 2nd ed. Weinheim, Germany: Wiley-VCH Verlag GmbH & Co.; 2006. [Google Scholar]
  • 44.Bowen TC, Noble RD, Falconer JL. Review: Fundamentals and applications of pervaporation through zeolite membranes. J Membr Sci. 2004;245:1–33. [Google Scholar]
  • 45.Vane LM. Pervaporation and Vapor Permeation Tutorial: Membrane Processes for the Selective Separation of Liquid and Vapor Mixtures. Sep Sci Technol. 2013;48(3):429–37. [Google Scholar]
  • 46.Offeman RD, Ludvik CN. A novel method to fabricate high permeance, high selectivity thin-film composite membranes. J Membr Sci. 2011;380:163–70. [Google Scholar]
  • 47.Vane LM, Namboodiri VV, Bowen TC. Hydrophobic zeolite-silicone rubber mixed matrix membranes for ethanol-water separation: Effect of zeolite and silicone component selection on pervaporation performance. J Membr Sci. 2008;308:230–41. [Google Scholar]
  • 48.Binning R, Lee R, Jennings J, Martin E. Separation of Liquid Mixtures by Permeation. IndEngChem. 1961;53(1):45–50. [Google Scholar]
  • 49.Wijmans JG, Baker RW. The solution-diffusion model: a review. J Membr Sci. 1995;107(1–2):1–21. [Google Scholar]
  • 50.Baker RW, Wijmans JG, Athayde AL, Daniels R, Ly JH, Le M. The Effect of Concentration Polarization on the Separation of Volatile Organic Compounds from Water by Pervaporation. J Membr Sci. 1997;137(1–2):159–72. [Google Scholar]
  • 51.Yeom CK, Huang RYM, editors. Pervaporation separation of aqueous mixtures using crosslinked poly(vinyl alcohol) (PVA): II. Permeation of ethanol-water mixtures. Proceedings of Fourth International Conference on Pervaporation Processes in the Chemical Industry; 1989; Fort Lauderdale, Florida: Bakish Materials Corporation. [Google Scholar]
  • 52.Cen Y, Wesslein M, Lichtenthaler RN, editors. Pervaporation of liquid mixtures through PVA membranes: New experimental results Proceedings of Fourth International Conference on Pervaporation Processes in the Chemical Industry; 1989; Fort Lauderdale, Florida: Bakish Materials Corporation. [Google Scholar]
  • 53.Wesslein M, Heintz A, Lichtenthaler RN. Pervaporation of liquid mixtures through poly(vinyl alcohol) (pva) membranes. I. study of water containing binary systems with complete and partial miscibility. J Membr Sci. 1990;51(1–2):169–79. [Google Scholar]
  • 54.Schehlmann MS, Wiedemann E, Lichtenthaler RN. Pervaporation and Vapor Permeation at the Azeotropic Point or in the Vicinity of the LLE Boundary Phases of Organic/Aqueous Mixtures. J Membr Sci. 1995;107(3):277–82. [Google Scholar]
  • 55.Will B, Lichtenthaler RN. Comparison of the Separation of Mixtures by Vapor Permeation and by Pervaporation Using PVA Composite Membranes. I. Binary Alcohol-Water Systems. J Membr Sci. 1992;68(1–2):119–25. [Google Scholar]
  • 56.Will B, Lichtenthaler RN. Comparison of the separation of mixtures by vapor permeation and by pervaporation using PVA composite membranes. II. The binary systems ammonia-water, methylamine-water, 1-propanol-methanol and the ternary system 1-propanol-methanol-water. J Membr Sci. 1992;68(1–2):127–31. [Google Scholar]
  • 57.Brüschke H, inventor; GFT Ingenieurburo fur Industrieanlagenbau (DE), assignee. Multi-layer membrane and the use thereof for the separation of liquid mixtures according to the pervaporation process. US Patent 4,755,299 1988.
  • 58.Brüschke H, inventor; GFT Ingenieurburo Fur Industrieanlagenbau, assignee. Multi-layer membrane and the use thereof for the separation of liquid mixtures according to the pervaporation process. U.S. Patent 4,915,834 1990.
  • 59.Fleming HL. Membrane pervaporation: Separation of organic/aqueous mixtures. Sep Sci Technol. 1990;25(13):1239–55. [Google Scholar]
  • 60.Pasternak M, Bartels CR, Reale J, Jr., inventors; Texaco Inc., assignee. Treatment of compositions containing water and organic oxygenates. U.S. Patent 4,910,344 1990.
  • 61.Gref R, Nguyen QT, Schaetzel P, Néel J. Transport properties of poly(vinyl alcohol) membranes of different degrees of crystallinity. I. Pervaporation results. J Appl Polym Sci. 1993;49(2):209–18. [Google Scholar]
  • 62.Néel JML, Nguyen QT, Brüschke H, inventors Composition membrane for separating water from fluids containing organic components by means of pervaporation. U.S. Patent 5,334,314 1994.
  • 63.Feng X, Huang RYM. Liquid separation by membrane pervaporation: a review. Ind Eng Chem Res. 1997;36:1048–66. [Google Scholar]
  • 64.Semenova SI, Ohya H, Soontarapa K. Hydrophilic Membranes for Pervaporation: An Analytical Review. Desalination. 1997;110:251–86. [Google Scholar]
  • 65.Jonquières A, Clément R, Lochon P, Néel J, Dresch M, Chretien B. Industrial state-of-the-art of pervaporation and vapour permeation in the western countries. J Membr Sci. 2002;206(1–2):87–117. [Google Scholar]
  • 66.Vane LM. A review of pervaporation for product recovery from biomass fermentation processes. J Chem Tech Biotechnol. 2005;80:603–29. [Google Scholar]
  • 67.Brüschke HEA, Marggraff F-K, Schäfer W, inventors; Sulzer Chemtech Gmbh - Membrantechnik, assignee. Composite membrane has pre-crosslinked PVA separation layer for separation of water from mixtures with organic compounds at up to specified temperature, obtained by post-crosslinking without addition of acids. German Patent DE102004060857A1 2006.
  • 68.Chapman PD, Oliveira T, Livingston AG, Li K. Membranes for the dehydration of solvents by pervaporation. J Membr Sci. 2008;318:5–37. [Google Scholar]
  • 69.Bolto B, Tran T, Hoang M, Xie Z. Crosslinked poly(vinyl alcohol) membranes. Progress in Polymer Science. 2009;34:969–81. [Google Scholar]
  • 70.Lipnizki F Membrane process opportunities and challenges in the bioethanol industry. Desalination. 2010;250:1067–9. [Google Scholar]
  • 71.Bolto B, Hoang M, Xie Z. A review of membrane selection for the dehydration of aqueous ethanol by pervaporation. Chem Eng Process: Process Intensification. 2011;50:227–35. [Google Scholar]
  • 72.Roizard D, Favre E. Trends in design and preparation of polymeric membranes for pervaporation In: Buonomenna MG, Golemme G, editors. Advanced Materials for Membrane Preparation: Bentham Science; 2012. p. 163–204. [Google Scholar]
  • 73.Bukusoglu E, Kalipcilar H, Yilmaz L. Dehydration of Industrial By-Product Solutions for Recycling via Pervaporation/Adsorption Hybrid Process. Ind Eng Chem Res. 2018;57(6):2277–86. [Google Scholar]
  • 74.Wong D, Parasrampuria J. Polyvinyl Alcohol. Analytical Profiles of Drug Substances and Excipients. 1996;24:397–441. [Google Scholar]
  • 75.Gohil JM, Bhattacharya A, Ray P. Studies on the crosslinking of poly(vinyl alcohol). Journal of Polymer Research. 2006;13(2):161–9. [Google Scholar]
  • 76.Jian S, Xiao Ming S. Crosslinked PVA-PS thin-film composite membrane for reverse osmosis. Desalination. 1987;62:395–403. [Google Scholar]
  • 77.Lang K, Sourirajan S, Matsuura T, Chowdhury G. A study on the preparation of polyvinyl alcohol thin-film composite membranes and reverse osmosis testing. Desalination. 1996;104(3):185–96. [Google Scholar]
  • 78.Xu S, Shen L, Li C, Wang Y. Properties and pervaporation performance of poly(vinyl alcohol) membranes crosslinked with various dianhydrides. J Appl Polym Sci. 2018;135(17):46159. [Google Scholar]
  • 79.Tang C, Saquing CD, Harding JR, Khan SA. In Situ Cross-Linking of Electrospun Poly(vinyl alcohol) Nanofibers. Macromolecules. 2009;43(2):630–7. [Google Scholar]
  • 80.Cordes EH, Bull HG. Mechanism and Catalysis for Hydrolysis of Acetals, Ketals, and Ortho Esters. Chemical Reviews. 1974;74(5):581–603. [Google Scholar]
  • 81.AG D Membrane data sheet: PERVAP™ polymeric membranes. AG DeltaMem; 2016. [Google Scholar]
  • 82.DeltaMem AG. Pervaporation (PV) and Vapour Permeation (VP) 2018. [Available from: http://www.deltamem.ch/ and https://www.deltamem.ch/pervaporation.html].
  • 83.The Folex Group: Membrane Separation Technology 2018. [Available from: http://www.folex.com/8-1-Special-Products.html]. [Google Scholar]
  • 84.Huang Y, Ly J, Aldajani T, Baker RW, inventors Liquid-phase and vapor-phase dehydration of organic/water solutions. US Patent 8,002,874 2011.
  • 85.Yeom CK, Kazi M, Baig FU. Simulation and Process Design of Pervaporation Plate-and-Frame Modules for Dehydration of Organic solvents. Membrane Journal. 2002;12(4):226–39. [Google Scholar]
  • 86.Petro Sep Corporation 2018. [Available from: http://petrosep.com/].
  • 87.KmX Membrane Technologies Corp. 2018. [Available from: http://kmxcorporation.wixsite.com/kmx-corporation].
  • 88.Baig FU. Pervaporation Advanced Membrane Technology and Applications: John Wiley & Sons, Inc.; 2008. p. 469–88. [Google Scholar]
  • 89.Kober PA. Pervaporation, perstillation and percrystallization. 1. Journal of the American Chemical Society. 1917;39(5):944–8. [Google Scholar]
  • 90.Yave W From collodion bag to skid-mounted pervaporation unit: 100th anniversary. Journal of Membrane Science and Research. 2017;3(4):263–4. [Google Scholar]
  • 91.Friesen DT, McCray SB, Newbold DD, Ray RJ, inventors; Bend Research, Inc., assignee. Solvent resistant hollow fiber vapor permeation membranes and modules. U.S. Patent 5,753,008 1998.
  • 92.Ong YK, Chung TS, inventors; National University of Singapore, assignee. Hollow fiber membrane for dehydration of organic solvents via pervaporation process and a method of fabricating the same. U.S. Patent Application US2015/0251140A1 2015.
  • 93.Huang Y, Baker RW, Wijmans JG. Perfluoro–coated hydrophilic membranes with improved selectivity. Ind Eng Chem Res. 2013;52:1141–9. [Google Scholar]
  • 94.Changluo Z, Moe L, Wei X. Separation of ethanol-water mixtures by pervaporation-membrane separation process. Desalination. 1987;62:299–313. [Google Scholar]
  • 95.Mulder MHV, Smolders CA. On the mechanism of separation of ethanol/water mixtures by pervaporation. I. Calculations of concentration profiles. J Membr Sci. 1984;17(3):289–307. [Google Scholar]
  • 96.Bell CM, Gerner FJ, Strathmann H. Selection of Polymers for Pervaporation Membranes. J Membr Sci. 1988;36(1):315–29. [Google Scholar]
  • 97.Reineke CE, Jagodzinski JA, Denslow KR. Highly water selective cellulosic polyelectrolyte membranes for the pervaporation of alcohol-water mixtures. J Membr Sci. 1987;32(2):207–21. [Google Scholar]
  • 98.Zuo J, Hua D, Maricar V, Ong YK, Chung T-S. Dehydration of industrial isopropanol (IPA) waste by pervaporation and vapor permeation membranes. J Appl Polym Sci. 2018;135(24):45086. [Google Scholar]
  • 99.Scholes CA, Stevens GW, Kentish SE. Membrane gas separation applications in natural gas processing. Fuel. 2012;96:15–28. [Google Scholar]
  • 100.Ji L, Shi B, Wang L. Pervaporation separation of ethanol/water mixture using modified zeolite filled PDMS membranes. J Appl Polym Sci. 2015;132(17):41897. [Google Scholar]
  • 101.Ohya H, Kudryavtsev VV, Semenova SI. Polyimide membranes : applications, fabrications, and properties. Australia: Gordon and Breach Publishers; 1996. [Google Scholar]
  • 102.Nakagawa K, Kusuki Y, Ninomiya K, editors. Separation of water-organic mixtures by vapor permeation through aromatic polyimide hollow fibers Proceedings of Fourth International Conference on Pervaporation Processes in the Chemical Industry; 1989; Fort Lauderdale, Florida: Bakish Materials Corporation. [Google Scholar]
  • 103.Okamoto K, Tanihara N, Watanabe H, Tanaka K, Kita H, Nakamura A, et al. Vapor Permeation and Pervaporation Separation of Water-Ethanol Mixtures Through Polyimide Membranes. J Membr Sci. 1992;68(1–2):53–63. [Google Scholar]
  • 104.Nakatani M, Sumiyama Y, Kusuki Y, inventors; Ube Industries, Ltd., assignee. Pervaporation method of selectively separating water from an organic material aqueous solution through aromatic imide polymer asymmetric membrane. US Patent 4,997,462. 1991.
  • 105.Makino H, Kusuki Y, Yoshida H, Nakamura A, inventors; Ube Industries, Ltd., assignee. Process for preparing aromatic polyimide semipermeable membranes. US Patent 4,378,324 1983.
  • 106.Japan Chemical Engineering & Machinery Co, Ltd.: Membrane separation type solvent dehydration system 2018. [Available from: http://www.nikkaki.co.jp/english/productguide/04.html].
  • 107.UBE Industries, Ltd.: UBE Organic Solvent Dehydration Membrane 2018. [Available from: https://www.ube.com/contents/en/chemical/separation/organic_solvent_dehydration_membrane.html].
  • 108.Hashimoto K, Ninomiya K, inventors; UBE Industries, LTD, assignee. Process and apparatus for separation of volatile components patent U.S. Patent No. 4,911,845 1990.
  • 109.Nakagawa K, Asakura Y, Yamamoto S, Ninomiya K, Kinouchi M, inventors; UBE Industries, Ltd., assignee. Method for dehydration and concentration of aqueous solution containing organic compound. US Patent 4,978,430 1990.
  • 110.Tanihara N, Nakanishi S, Yoshinaga T. Gas and Vapor Separation through Polyimide Membranes. Journal of the Japan Petroleum Institute. 2016;59(6):276–82. [Google Scholar]
  • 111.Tanihara N, Tanaka K, Kita H, Okamoto K-i, Nakamura A, Kusuki Y, et al. Vapor-Permeation Separation of Water-Ethanol Mixtures by Asymmetric Polyimide Hollow-Fiber Membrane Modules. Journal of Chemical Engineering of Japan. 1992;25(4):388–96. [Google Scholar]
  • 112.Roth T, Kreis P, Górak A. Process analysis and optimisation of hybrid processes for the dehydration of ethanol. Chemical Engineering Research and Design. 2013;91:1171–85. [Google Scholar]
  • 113.Roth T, Kreis P. Rate Based Modelling and Simulation Studies of Hybrid Processes consisting of Distillation, Vapour Permeation and Adsorption for the dehydration of ethanol. Computer Aided Chemical Engineering. 2009;26:815–9. [Google Scholar]
  • 114.Basu A. Membrane Startups Differentiate With Novel Materials in Crowded Municipal and Industrial Markets (Lux Research blog post) 2017. [Available from: http://blog.luxresearchinc.com/blog/2017/08/membrane-startups-differentiate-with-novel-materials-in-crowded-municipal-and-industrial-markets/].
  • 115.Plante P, de Caumia B, Roy C, Noel G, inventors; Parker Filtration BV (originally assigned to Vaperma Inc.), assignee. Ethanol processing with vapour separation membranes. U.S. Patent 8,128,826 2012.
  • 116.Cote P, Roy C, Bernier N, Schwartz M, Dodkin T, Bradt C. Field Demonstration of the Siftek™ Membrane for Ethanol Dewatering. 2007.
  • 117.Cranford R, Roy C, inventors; Vaperma Inc., assignee. Solvent resistant asymmetric integrally skinned membranes. US Patent 7,556,677 2009.
  • 118.Tanaka K, Okamoto K-I. Structure and transport properties of polyimides as materials for gas and vapor membrane separation In: Yampolskii Y, Pinnau I, Freeman BD, editors. Materials Science of Membranes for Gas and Vapor Separation. West Sussex, England: John Wiley & Sons, Ltd.; 2006. [Google Scholar]
  • 119.Khosravi T, Mosleh S, Bakhtiari O, Mohammadi T. Mixed matrix membranes of Matrimid 5218 loaded with zeolite 4A for pervaporation separation of water-isopropanol mixtures. Chemical Engineering Research and Design. 2012;90:2353–63. [Google Scholar]
  • 120.Jiang LY, Chung TS. Homogeneous polyimide/cyclodextrin composite membranes for pervaporation dehydration of isopropanol. J Membr Sci. 2010;346(1):45–58. [Google Scholar]
  • 121.Qiu W, Kosuri M, Zhou F, Koros WJ. Dehydration of ethanol-water mixtures using asymmetric hollow fiber membranes from commercial polyimides. J Membr Sci. 2009;327:96–103. [Google Scholar]
  • 122.Kreiter R, Wolfs DP, Engelen CWR, van Veen HM, Vente JF. High-temperature pervaporation performance of ceramic-supported polyimide membranes in the dehydration of alcohols. J Membr Sci. 2008;319(1–2):126–32. [Google Scholar]
  • 123.Mangindaan DW, Min Shi G, Chung T-S. Pervaporation dehydration of acetone using P84 co-polyimide flat sheet membranes modified by vapor phase crosslinking. J Membr Sci. 2014;458:76–85. [Google Scholar]
  • 124.Hua D, Ong YK, Wang Y, Yang T, Chung T-S. ZIF-90/P84 mixed matrix membranes for pervaporation dehydration of isopropanol. J Membr Sci. 2014;453(0):155–67. [Google Scholar]
  • 125.Liu R, Qiao X, Chung TS. The development of high performance P84 co-polyimide hollow fibers for pervaporation dehydration of isopropanol. Chem Eng Sci. 2005;60:6674–86. [Google Scholar]
  • 126.Shi GM, Wang Y, Chung T-S. Dual-layer PBI/P84 hollow fibers for pervaporation dehydration of acetone. AIChE Journal. 2012;58(4):1133–45. [Google Scholar]
  • 127.Kang Ong Y, Chung T-S. High performance dual-layer hollow fiber fabricated via novel immiscibility induced phase separation (I2PS) process for dehydration of ethanol. J Membr Sci. 2012;421–422:271–82. [Google Scholar]
  • 128.Kang Ong Y, Wang H, Chung T-S. A prospective study on the application of thermally rearranged acetate containing polyimide membranes in dehydration of biofuels via pervaporation. Chem Eng Sci. 2012;79:41–53. [Google Scholar]
  • 129.Freeman BD, Paul DR, Czenkusch K, Ribeiro Jr CP, Ba C, inventors; Board of Regents, The University of Texas System, assignee. Thermally Rearranged (TR) Polymers as Membranes for Ethanol Dehydration US Patent Application 20120305484 A1 2012.
  • 130.Czenkusch K Evaluation of Transport and Transport Stability in Glassy Polymer Membranes: The University of Texas at Austin; 2014. [Google Scholar]
  • 131.Scholes CA, Freeman BD, Kentish SE. Water vapor permeability and competitive sorption in thermally rearranged (TR) membranes. J Membr Sci. 2014;470(0):132–7. [Google Scholar]
  • 132.O’Brien KC, Krishnan G, Berchtold KA, Blum S, Callahan R, Johnson W, et al. Towards a pilot-scale membrane system for pre-combustion CO2 separation. Energy Procedia. 2009;1(1):287–94. [Google Scholar]
  • 133.Jiraratananon R, Chanachai A, Huang RYM. Pervaporation dehydration of ethanol–water mixtures with chitosan/hydroxyethylcellulose (CS/HEC) composite membranes: II. Analysis of mass transport. J Membr Sci. 2002;199(1–2):211–22. [Google Scholar]
  • 134.Huang RYM, Pal R, Moon GY. Crosslinked Chitosan Composite Membrane for the Pervaporation Dehydration of Alcohol Mixtures and Enhancement of Structural Stability of Chitosan/Polysulfone Composite Membranes. J Membr Sci. 1999;160(1):17–30. [Google Scholar]
  • 135.Shao P, Huang RYM. Polymeric membrane pervaporation. J Membr Sci. 2007;287(2):162–79. [Google Scholar]
  • 136.Wang X-S, An Q-F, Zhao Q, Lee K-R, Qian J-W, Gao C-J. Homogenous polyelectrolyte complex membranes incorporated with strong ion-pairs with high pervaporation performance for dehydration of ethanol. J Membr Sci. 2013;435:71–9. [Google Scholar]
  • 137.Childs RF, Yu J, inventors; McMaster University, assignee. Pervaporation composite membranes. US Patent 7,604,746 B2 2009.
  • 138.Liu Y, Zhu M, Zhao Q, An Q, Qian J, Lee K, et al. The chemical crosslinking of polyelectrolyte complex colloidal particles and the pervaporation performance of their membranes. J Membr Sci. 2011;385–386(0):132–40. [Google Scholar]
  • 139.Zhu M, Qian J, Zhao Q, An Q, Song Y, Zheng Q. Polyelectrolyte complex (PEC) modified by poly (vinyl alcohol) and their blend membranes for pervaporation dehydration. J Membr Sci. 2011;378:233–42. [Google Scholar]
  • 140.Urtiaga AM, Casado C, Aragoza C, Ortiz I. Dehydration of industrial ketonic effluents by pervaporation. Comparative behavior of ceramic and polymeric membranes. Sep Sci Technol. 2003(38):3473–91. [Google Scholar]
  • 141.Zhao Q, Qian J, An Q, Sun Z. Layer-by-layer self-assembly of polyelectrolyte complexes and their multilayer films for pervaporation dehydration of isopropanol. J Membr Sci. 2010;346:335–43. [Google Scholar]
  • 142.Zhang Y, Rhim JW, Feng X. Improving the stability of layer-by-layer self-assembled membranes for dehydration of alcohol and diol. J Membr Sci. 2013;444:22–31. [Google Scholar]
  • 143.Das P, Ray SK. Synthesis and characterization of biopolymer based mixed matrix membranes for pervaporative dehydration. Carbohydrate Polymers. 2014;103:274–84. [DOI] [PubMed] [Google Scholar]
  • 144.Das P, Ray SK. Analysis of sorption and permeation of acetic acid-water mixtures through unfilled and filled blend membranes. Sep Purif Technol. 2013;116:433–47. [Google Scholar]
  • 145.Bartels CR, inventor; Texaco, Inc., assignee. Separation of compositions containing water and organic oxygenates. U.S. Patent 4,971,699 1990.
  • 146.Gudeman LF, Peppas NA. pH-sensitive membranes from poly(vinyl alcohol)/poly(acrylic acid) interpenetrating networks. J Membr Sci. 1995;107(3):239–48. [Google Scholar]
  • 147.Ping Z, Nguyen Q-T, Essamri A, Néel J. High-performance membranes for pervaporation II. Crosslinked poly(vinyl alcohol)–poly(acrylic acid) blends. Polymers for Advanced Technologies. 1994;5(6):320–6. [Google Scholar]
  • 148.Ruckenstein E, Liang L. Poly(acrylic acid)-poly(vinyl alcohol) semi- and interpenetrating polymer network pervaporation membranes. J Appl Polym Sci. 1996;62(7):973–87. [Google Scholar]
  • 149.Anjali Devi D, Smitha B, Sridhar S, Aminabhavi TM. Pervaporation separation of dimethylformamide/water mixtures through poly(vinyl alcohol)/poly(acrylic acid) blend membranes. Sep Purif Technol. 2006;51(1):104–11. [Google Scholar]
  • 150.Bano S, Mahmood A, Lee K-H. Vapor Permeation Separation of Methanol-Water Mixtures: Effect of Experimental Conditions. Ind Eng Chem Res. 2013;52:10450–9. [Google Scholar]
  • 151.Namboodiri VV, Ponangi RP, Vane LM. A novel hydrophilic polymer membrane for the dehydration of organic solvents. EurPolymJ. 2006;42:3390–3. [Google Scholar]
  • 152.Namboodiri VV, Vane LM. High permeability membranes for the dehydration of low water content ethanol by pervaporation. J Membr Sci. 2007;306(1–2):209–15. [Google Scholar]
  • 153.Tang J Separation of various organic-organic and aqueous-organic solutions via pervaporation: New Jersey Institute of Technology; 2013. [Google Scholar]
  • 154.Jalal TA, Bettahalli NMS, Le NL, Nunes SP. Hydrophobic Hyflon AD/Poly(vinylidene fluoride) Membranes for Butanol Dehydration via Pervaporation. Ind Eng Chem Res. 2015;54(44):11180–7. [Google Scholar]
  • 155.Arcella V, Ghielmi A, Tommasi G. High performance perfluoropolymer films and membranes. AnnNYAcadSci. 2003;984:226–44. [DOI] [PubMed] [Google Scholar]
  • 156.Merkel TC, Zhang H, He Z, Wijmans JG, Okamoto Y, inventors; Membrane Technology and Research, Inc. New York University, assignee. Fluid Separation Processes Using Membranes Based on Fluorinated and Perfluorinated Polymers. U.S. Patent 9,975,084 2018.
  • 157.Smuleac V, Wu J, Nemser S, Majumdar S, Bhattacharyya D. Novel Perfluorinated Polymer-Based Pervaporation Membranes for Separation of Solvent/Water Mixtures. J Membr Sci. 2010;352:41–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Majumdar S, Stookey D, Nemser S, Campos D, Pennisi K. Perfluoropolymer Membranes for Dehydration of Ethanol, presented at AIChE Annual Meeting, Philadelphia, PA: 2008. [Google Scholar]
  • 159.Huang Y, Ly J, Nguyen D, Baker RW. Ethanol Dehydration Using Hydrophobic and Hydrophilic Polymer Membranes. Ind Eng Chem Res. 2010;49(23):12067–73. [Google Scholar]
  • 160.Scholes CA, Kanehashi S, Stevens GW, Kentish SE. Water permeability and competitive permeation with CO2 and CH4 in perfluorinated polymeric membranes. Sep Purif Technol. 2015;147:203–9. [Google Scholar]
  • 161.Campos D, Knopf J, Majumdar S, Nemser S. THF Dehydration by Pervaporation with Perfluorinated Composite Membrane, presented at North American Membrane Society annual meeting, Las Vegas, NV: 2011. [Google Scholar]
  • 162.Tang J, Sirkar KK, Majumdar S. Pervaporative dehydration of concentrated aqueous solutions of selected polar organics by a perfluoropolymer membrane. Sep Purif Technol. 2017;175:122–9. [Google Scholar]
  • 163.Tang J, Sirkar KK. Perfluoropolymer membrane behaves like a zeolite membrane in dehydration of aprotic solvents. J Membr Sci. 2012;421–422:211–6. [Google Scholar]
  • 164.Kiyono M, Williams PJ, Koros WJ. Effect of pyrolysis atmosphere on separation performance of carbon molecular sieve membranes. J Membr Sci. 2010;359(1):2–10. [Google Scholar]
  • 165.Uragami T 2.11 - Selective Membranes for Purification and Separation of Organic Liquid Mixtures Comprehensive Membrane Science and Engineering. Oxford: Elsevier; 2010. p. 273–324. [Google Scholar]
  • 166.Liao K-S, Fu Y-J, Hu C-C, Chen J-T, Lin D-W, Lee K-R, et al. Microstructure of carbon molecular sieve membranes and their application to separation of aqueous bioethanol. Carbon. 2012;50(11):4220–7. [Google Scholar]
  • 167.Tin PS, Lin HY, Ong RC, Chung T-S. Carbon molecular sieve membranes for biofuel separation. Carbon. 2011;49:369–75. [Google Scholar]
  • 168.Ichikawa A, Suzuki K, Kinoshita N, Isoda Y, Kimata T, inventors; NGK Insulators, Ltd., assignee. Carbon membrane and method for pervaporation separation. U.S. Patent 8,945,390 2015.
  • 169.Media and Process Technology Inc.: Water Permeation Membranes for Dehydration via Pervaporation Applications 2018. [Available from: http://www.mediaandprocess.com/products/products03.html]. [Google Scholar]
  • 170.Park HB, Kamcev J, Robeson LM, Elimelech M, Freeman BD. Maximizing the right stuff: The trade-off between membrane permeability and selectivity. Science. 2017;356(6343):1137. [DOI] [PubMed] [Google Scholar]
  • 171.Sun P, Wang K, Zhu H. Recent Developments in Graphene-Based Membranes: Structure, Mass-Transport Mechanism and Potential Applications. Advanced Materials. 2016;28(12):2287–310. [DOI] [PubMed] [Google Scholar]
  • 172.Ma J, Ping D, Dong X. Recent Developments of Graphene Oxide-Based Membranes: A Review. Membranes. 2017;7(3):52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Van Gestel T, Barthel J. New types of graphene-based membranes with molecular sieve properties for He, H2 and H2O. J Membr Sci. 2018;554:378–84. [Google Scholar]
  • 174.Paulauskas M Pervaporation Using Graphene Oxide Membranes: The University of Leeds; 2015. [Google Scholar]
  • 175.Tang YP, Paul DR, Chung TS. Free-standing graphene oxide thin films assembled by a pressurized ultrafiltration method for dehydration of ethanol. J Membr Sci. 2014;458:199–208. [Google Scholar]
  • 176.Huang K, Liu G, Jin W. Vapor transport in graphene oxide laminates and their application in pervaporation. Current Opinion in Chemical Engineering. 2017;16:56–64. [Google Scholar]
  • 177.Huang K, Liu G, Lou Y, Dong Z, Shen J, Jin W. A Graphene Oxide Membrane with Highly Selective Molecular Separation of Aqueous Organic Solution. Angewandte Chemie International Edition. 2014;53(27):6929–32. [DOI] [PubMed] [Google Scholar]
  • 178.Lou Y, Liu G, Liu S, Shen J, Jin W. A facile way to prepare ceramic-supported graphene oxide composite membrane via silane-graft modification. Applied Surface Science. 2014;307:631–7. [Google Scholar]
  • 179.Hung W-S, An Q-F, De Guzman M, Lin H-Y, Huang S-H, Liu W-R, et al. Pressure-assisted self-assembly technique for fabricating composite membranes consisting of highly ordered selective laminate layers of amphiphilic graphene oxide. Carbon. 2014;68:670–7. [Google Scholar]
  • 180.Nair R, Budd P, Geim A, inventors; The University of Manchester, assignee. Separation of water using a membrane. U.S. Patent 9,844,758 2017.
  • 181.Chu B, Hsiao BS, Mahajan D, Yeh T-M, inventors; The Research Foundation For The State University of New York, assignee. Graphene oxide-based composite membranes. U.S. Patent 9,353,037 2016.
  • 182.McKeown NB, Budd PM, Msayib KJ, Ghanem BS, Kingston HJ, Tattershall CE, et al. Polymers of intrinsic microporosity (PIMs): bridging the void between microporous and polymeric materials. Chemistry. 2005;11:2610–20. [DOI] [PubMed] [Google Scholar]
  • 183.McKeown NB, Budd PM, Msayib K, Ghanem B, inventors; The University of Manchester, assignee. Microporous polymer material. U.S. Patent 8,056,732 2011.
  • 184.Rogers CE, Stannett V, Szwarc M. Permeability Valves. Permeability of Gases and Vapors through Composite Membranes. Ind Eng Chem. 1957;49(11):1933–6. [Google Scholar]
  • 185.Kuriyama T, Nobutoki M, Nakanishi M. Permeability of Double-Layer Films III. JPharmSci. 1970;59(10):1412–9. [DOI] [PubMed] [Google Scholar]
  • 186.Huang Y, Baker RW, Aldajani T, Ly J, inventors; Membrane Technology and Research, Inc., assignee. Dehydration processes using membranes with hydrophobic coating. US Patent 9,061,252 2015.
  • 187.Hua D, Chung T-S, Shi GM, Fang C. Teflon AF2400/Ultem composite hollow fiber membranes for alcohol dehydration by high-temperature vapor permeation. AIChE Journal. 2016;62(5):1747–57. [Google Scholar]
  • 188.Hua D. Design and fabrication of polymer based membranes for alcohol dehyration via pervaporation or vapor permeation (Ph.D. thesis): National University of Singapore; 2016. [Google Scholar]
  • 189.Wee S-L, Tye C-T, Bhatia S. Membrane separation process--Pervaporation through zeolite membrane. Sep Purif Technol. 2008;63(3):500–16. [Google Scholar]
  • 190.Liu W, Li XS, Canfield NL, inventors; Battelle Memorial Institute, assignee. Thin-sheet zeolite membrane and methods for making the same. U.S. Patent Application 2015/0265975 2015.
  • 191.Lin YS. Microporous and dense inorganic membranes: current status and prospective. Sep Purif Technol. 2001;25(1–3):39–55. [Google Scholar]
  • 192.Jha B, Singh DN. Chapter 2: Basics of Zeolites In: Jha B, Singh DN, editors. Fly Ash Zeolites Advanced Structured Materials: Springer Singapore; 2016. p. 5–31. [Google Scholar]
  • 193.Cui Y, Kita H, Okamoto KI. Zeolite T membrane: preparation, characterization, pervaporation of water/organic liquid mixtures and acid stability. J Membr Sci. 2004;236(1–2):17–27. [Google Scholar]
  • 194.Pera-Titus M, Llorens J. Zeolite A Membranes for Solvent Dehydration: An Overview and Review of Own Results In: Wong TW, editor. Handbook of Zeolites: Structure, Properties and Applications. Materials Science and Technologies Series. New York: Nova Science Publishers, Inc; 2009. p. 523–54. [Google Scholar]
  • 195.van Leeuwen ME. Derivation of Stockmayer potential parameters for polar fluids. Fluid Phase Equilibria. 1994;99:1–18. [Google Scholar]
  • 196.Van der Bruggen B, Schaep J, Wilms D, Vandecasteele C. Influence of molecular size, polarity and charge on the retention of organic molecules by nanofiltration. J Membr Sci. 1999;156(1):29–41. [Google Scholar]
  • 197.Yamamoto T, Kim YH, Kim BC, Endo A, Thongprachan N, Ohmori T. Adsorption characteristics of zeolites for dehydration of ethanol: Evaluation of diffusivity of water in porous structure. Chem Eng J. 2012;181–182(0):443–8. [Google Scholar]
  • 198.Kondo M, Kita H. Permeation mechanism through zeolite NaA and T-type membranes for practical dehydration of organic solvents. J Membr Sci. 2010;361:223–31. [Google Scholar]
  • 199.Kita H, Horii K, Ohtoshi Y, Tanaka K, Okamoto KI. Synthesis of a Zeolite NaA Membrane for Pervaporation of Water/Organic Liquid Mixtures. Journal of Materials Science Letters. 1995;14:206–8. [Google Scholar]
  • 200.Okamoto KI, Kita H, Horii K, Tanaka K, Kondo M. Zeolite NaA membrane: Preparation, single-gas permeation, and pervaporation and vapor permeation of water/organic liquid mixtures. Ind Eng Chem Res. 2001;40:163–75. [Google Scholar]
  • 201.Pera-Titus M. Preparation, characterization and modeling of zeolite NaA membranes for the pervaporation dehydration of alcohol mixtures (Ph.D. thesis): University of Barcelona; 2006. [Google Scholar]
  • 202.Li Y, Zhou H, Zhu G, Liu J, Yang W. Hydrothermal stability of LTA zeolite membranes in pervaporation. J Membr Sci. 2007;297(1–2):10–5. [Google Scholar]
  • 203.Rangnekar N, Mittal N, Elyassi B, Caro J, Tsapatsis M. Zeolite membranes - a review and comparison with MOFs. Chemical Society Reviews. 2015;44(20):7128–54. [DOI] [PubMed] [Google Scholar]
  • 204.Hasegawa Y, Nagase T, Kiyozumi Y, Hanaoka T, Mizukami F. Influence of acid on the permeation properties of NaA-type zeolite membranes. J Membr Sci. 2010;349(1–2):189–94. [Google Scholar]
  • 205.Yu C, Liu Y, Chen G, Gu X, Xing W. Pretreatment of Isopropanol Solution from Pharmaceutical Industry and Pervaporation Dehydration by NaA Zeolite Membranes. Chinese Journal of Chemical Engineering. 2011;19(6):904–10. [Google Scholar]
  • 206.Cho CH, Oh KY, Yeo JG, Kim SK, Han MH, Lee YM. Synthesis, Ethanol Dehydration and Thermal Stability of NaA Zeolite/Alumina Composite Membranes with Narrow Non-zeolitic Pores and Thin Intermediate Layer. J Membr Sci. 2010;364:138–48. [Google Scholar]
  • 207.Kallus S, Condre JM, Hahn A, Golemme G, Algieri C, Dieudonne P, et al. Colloidal zeolites and zeolite membranes. Journal of Materials Chemistry. 2002;12:1–8. [Google Scholar]
  • 208.Choi J, Jeong H-K, Snyder MA, Stoeger JA, Masel RI, Tsapatsis M. Grain Boundary Defect Elimination in a Zeolite Membrane by Rapid Thermal Processing. Science. 2009;325(5940):590–3. [DOI] [PubMed] [Google Scholar]
  • 209.Lin YS, Duke MC. Recent progress in polycrystalline zeolite membrane research. Current Opinion in Chemical Engineering. 2013;2:209–16. [Google Scholar]
  • 210.Ye P, Zhang Y, Wu H, Gu X. Mass transfer simulation on pervaporation dehydration of ethanol through hollow fiber NaA zeolite membranes. AIChE Journal. 2016;62(7):2468–78. [Google Scholar]
  • 211.Wang J, Ye P, Gao X, Zhang Y, Gu X. Modeling investigation of geometric size effect on pervaporation dehydration through scaled-up hollow fiber NaA zeolite membranes. Chinese Journal of Chemical Engineering. 2018;26(7):1477–84. [Google Scholar]
  • 212.Ji M, Gao X, Wang X, Zhang Y, Jiang J, Gu X. An ensemble synthesis strategy for fabrication of hollow fiber T-type zeolite membrane modules. J Membr Sci. 2018;563:460–9. [Google Scholar]
  • 213.Jiang J, Peng L, Wang X, Qiu H, Ji M, Gu X. Effect of Si/Al ratio in the framework on the pervaporation properties of hollow fiber CHA zeolite membranes. Microporous and Mesoporous Materials. 2019;273:196–202. [Google Scholar]
  • 214.Kita H, Horii K, Tanaka K, Okamoto K-I, Miyake N, Kondo M. Pervaporation of water/organic liquid mixtures using a zeolite NaA membrane In: Bakish R, editor. Proceedings of Seventh International Conference on Pervaporation Processes in the Chemical Industry; Reno, Nevada USA: Bakish Materials Corporation; 1995. p. 364–73. [Google Scholar]
  • 215.Okamoto K, Kita H, Kondo M, Miyake N, Matsuo Y, inventors; Mitsui Engineering & Shipbuilding Co., Ltd., assignee. Membrane for liquid mixture separation. U.S. Patent 5,554,286 1996.
  • 216.Membrane Zeolite (*PV/VP Separation): Mitsui Engineering & Shipbuilding Co., Ltd; 2018. [Available from: https://www.mes.co.jp/english/business/environment/environ_plant/detail764.html].
  • 217.Guerra K, Pellegrino J. Development of a Techno-Economic Model to Compare Ceramic and Polymeric Membranes. Sep Sci Technol. 2013;48:51–65. [Google Scholar]
  • 218.Wenten IG, Dharmawijaya PT, Aryanti PTP, Mukti RR, Khoiruddin. LTA zeolite membranes: current progress and challenges in pervaporation. RSC Advances. 2017;7(47):29520–39. [Google Scholar]
  • 219.Holtbruegge J, Kuhlmann H, Lutze P. Process Analysis and Economic Optimization of Intensified Process Alternatives for Simultaneous Industrial Scale Production of Dimethyl Carbonate and Propylene Glycol. Chemical Engineering Research and Design. 2015;93:411–31. [Google Scholar]
  • 220.Vane LM, Alvarez FR. Effect of membrane and process characteristics on cost and energy usage for separating alcohol-water mixtures using hybrid vapor stripping-vapor permeation process. J Chem Technol Biotechnol. 2015;90:1380–90. [Google Scholar]
  • 221.Billard P, Kind M. Performance of PVA- and zeolite-membranes in selective drying processes. Chem Eng Process. 2003;42(1):23–8. [Google Scholar]
  • 222.Wang R, Ma N, Yan Y, Wang Z. Ultrasonic-assisted fabrication of high flux T-type zeolite membranes on alumina hollow fibers. J Membr Sci. 2018;548:676–84. [Google Scholar]
  • 223.Chen X, Wang J, Yin D, Yang J, Lu J, Zhang Y, et al. High-performance zeolite T membrane for dehydration of organics by a new varying temperature hot-dip coating method. AIChE Journal. 2013;59(3):936–47. [Google Scholar]
  • 224.Wang X, Jiang J, Liu D, Xue Y, Zhang C, Gu X. Evaluation of hollow fiber T-type zeolite membrane modules for ethanol dehydration. Chinese Journal of Chemical Engineering. 2017;25(5):581–6. [Google Scholar]
  • 225.Kondo M, Morigami Y, Okamoto K, Kita H, inventors; Mitsui Engineering & Shipbuilding Co., Ltd., assignee. Process for producing a membrane for separating a mixture. U.S. Patent 6,159,542 2000.
  • 226.Press Release: Mitsui Zosen Machinery & Service and Mitsubishi Chemical to Form Tie-up to Produce and Sell Zeolite Membranes 2016. [Available from: http://www.mes.co.jp/english/press/2016/20160714.html].
  • 227.Mitsubishi Chemical, ICM bring ZEBREX to US ethanol industry 2018. [Available from: http://ethanolproducer.com/articles/15071/mitsubishi-chemical-icm-bring-zebrex-to-us-ethanol-industry].
  • 228.Maeda K, Yamazaki K, Kakiuchi H, Tahara T, Takewaki T, inventors; Mitsubishi Chemical Corporation and Mitsubishi Chemical Engineering Corporation, assignee. Method for producing high-concentration alcohol. U.S. Patent Application 2017/0204030 2017.
  • 229.Sugita M, Takewaki T, Oshima K, Fujita N, inventors; Mitsubishi Chemical Corporation, assignee. Inorganic porous support-zeolite membrane composite, production method thereof, and separation method using the composite. U.S. Patent 8,376,148 2013.
  • 230.Hitachi Zosen Corporation: Hitz Dehydration System HDS® by Zeolite Membrane Element 2018. [Available from: http://www.hitachizosen.co.jp/english/products/products009.html and http://www.hitachizosen.co.jp/english/technology/hitz-tech/environment08.html].
  • 231.Imasaka S, Itakura M, Hasegawa Y, Sato K, inventors; Hitachi Zosen Corporation, assignee. Zeolite membrane, production method therefor, and separation method using same. U.S. Patent application 2017/0189862 2017.
  • 232.Imasaka S, Itakura M, Yano K, Fujita S, Okada M, Hasegawa Y, et al. Rapid preparation of high-silica CHA-type zeolite membranes and their separation properties. Sep Purif Technol. 2018;199:298–303. [Google Scholar]
  • 233.Baerlocher C, McCusker L. Database of Zeolite Structures [Available from: http://www.iza-structure.org/databases/].
  • 234.Denayer JFM, Devriese LI, Couck S, Martens J, Singh R, Webley PA, et al. Cage and Window Effects in the Adsorption of n-Alkanes on Chabazite and SAPO-34. The Journal of Physical Chemistry C. 2008;112(42):16593–9. [Google Scholar]
  • 235.Sato K, Sugimoto K, Shimotsuma N, Kikuchi T, Kyotani T, Kurata T. Development of practically available up-scaled high-silica CHA-type zeolite membranes for industrial purpose in dehydration of N-methyl pyrrolidone solution. J Membr Sci. 2012;409–410:82–95. [Google Scholar]
  • 236.Li X, Kita H, Zhu H, Zhang Z, Tanaka K, Okamoto K-i. Influence of the hydrothermal synthetic parameters on the pervaporative separation performances of CHA-type zeolite membranes. Microporous and Mesoporous Materials. 2011;143(2):270–6. [Google Scholar]
  • 237.Hasegawa Y, Abe C, Nishioka M, Sato K, Nagase T, Hanaoka T. Formation of high flux CHA-type zeolite membranes and their application to the dehydration of alcohol solutions. J Membr Sci. 2010;364(1):318–24. [Google Scholar]
  • 238.van Gemert RW, Cuperus FP. Newly Developed Ceramic Membranes for Dehydration and Separation of Organic Mixtures by Pervaporation. J Membr Sci. 1995;105(3):287–91. [Google Scholar]
  • 239.van Veen HM. Reduction of energy consumption in the process industry by pervaporation with inorganic membranes: techno-economical feasibility study. ECN; 2001. [Google Scholar]
  • 240.van Veen HM, van Delft YC, Engelen CWR, Pex PPAC. Dewatering of organics by pervaporation with silica membranes. Sep Purif Technol. 2001;22–23(1–711):361–6. [Google Scholar]
  • 241.Gallego-Lizon T, Ho YS, dos Santos LF. Comparative study of commercially available polymeric and microporous silica membranes for the dehydration of IPA/water mixtures by pervaporation/vapour permeation. Desalination. 2002;149(1–3):3–8. [Google Scholar]
  • 242.Castricum HL, Sah A, Kreiter R, Blank DHA, Vente JF, ten Elshof JE. Hydrothermally stable molecular separation membranes from organically linked silica. Journal of Materials Chemistry. 2008;18(18):2150–8. [Google Scholar]
  • 243.Paradis GG. Novel concepts for microporous hybrid silica membranes - Functionalization & pore size tuning: University of Twente; 2012. [Google Scholar]
  • 244.ten Hove M, Luiten-Olieman MWJ, Huiskes C, Nijmeijer A, Winnubst L. Hydrothermal stability of silica, hybrid silica and Zr-doped hybrid silica membranes. Sep Purif Technol. 2017;189:48–53. [Google Scholar]
  • 245.Kreiter R, Castricum HL, Vente JF, Ten Elshof JE, Rietkerk MDA, Veen HM, inventors; Stichting Energieonderzoek Centrum Nederland, assignee. Hybrid silica membrane for water removal from lower alcohols and hydrogen separation U.S. Patent 8,709,254 2014.
  • 246.Sah A, Castricum HL, Vente JF, Blank DHA, ten Elshof JE, inventors; Stichting Energieonderzoek Centrum Nederland, assignee. Microporous molecular separation membrane with high hydrothermal stability. U.S. Patent 8,277,661 2012.
  • 247.Agirre Arisketa I, Arias PL, Castricum HL, Creatore M, ten Elshof JE, Paradis GG, et al. Hybrid Organosilica Membranes and Processes: status and outlook. Sep Purif Technol. 2014;121:2–12. [Google Scholar]
  • 248.Castricum HL, Paradis GG, Mittelmeijer-Hazeleger MC, Bras W, Eeckhaut G, Vente JF, et al. Tuning the nanopore structure and separation behavior of hybrid organosilica membranes. Microporous and Mesoporous Materials. 2014;185:224–34. [Google Scholar]
  • 249.Paradis GG, Shanahan DP, Kreiter R, van Veen HM, Castricum HL, Nijmeijer A, et al. From hydrophilic to hydrophobic HybSi® membranes: A change of affinity and applicability. J Membr Sci. 2013;428:157–62. [Google Scholar]
  • 250.Wang J, Kanezashi M, Yoshioka T, Tsuru T. Effect of calcination temperature on the PV dehydration performance of alcohol aqueous solutions through BTESE-derived silica membranes. J Membr Sci. 2012;415–416:810–5. [Google Scholar]
  • 251.Dral AP, van Eck ERH, Winnubst L, ten Elshof JE. Micropore structure stabilization in organosilica membranes by gaseous catalyst post-treatment. J Membr Sci. 2018;548:157–64. [Google Scholar]
  • 252.Moriyama N, Nagasawa H, Kanezashi M, Tsuru T. Pervaporation dehydration of aqueous solutions of various types of molecules via organosilica membranes: Effect of membrane pore sizes and molecular sizes. Sep Purif Technol. 2018;207:108–15. [Google Scholar]
  • 253.PERVATECH BV Datasheet: Hybrid Silica HybSi® AR Membranes; 2018. [Google Scholar]
  • 254.PERVATECH BV 2018. [Available from: http://pervaporation-membranes.com/].
  • 255.Nagasawa H, Matsuda N, Kanezashi M, Yoshioka T, Tsuru T. Pervaporation and vapor permeation characteristics of BTESE-derived organosilica membranes and their long-term stability in a high-water-content IPA/water mixture. J Membr Sci. 2016;498:336–44. [Google Scholar]
  • 256.Kreiter R, Rietkerk MDA, Castricum HL, van Veen HM, ten Elshof JE, Vente JF. Stable Hybrid Silica Nanosieve Membranes for the Dehydration of Lower Alcohols. ChemSusChem. 2009;2(2):158–60. [DOI] [PubMed] [Google Scholar]
  • 257.Castricum HL, Kreiter R, van Veen HM, Blank DHA, Vente JF, ten Elshof JE. High-performance hybrid pervaporation membranes with superior hydrothermal and acid stability. J Membr Sci. 2008;324:111–8. [Google Scholar]
  • 258.van Veen HM, Rietkerk MDA, Shanahan DP, van Tuel MMA, Kreiter R, Castricum HL, et al. Pushing membrane stability boundaries with HybSi® pervaporation membranes. J Membr Sci. 2011;380:124–31. [Google Scholar]
  • 259.Itoh N, Ishida J, Kikuchi Y, Sato T, Hasegawa Y. Continuous dehydration of IPA–water mixture by vapor permeation using Y type zeolite membrane in a recycling system. Sep Purif Technol. 2015;147:346–52. [Google Scholar]
  • 260.Mizuno T, Katakura Y, inventors; Mitsubishi Chemical Corporation, assignee. Process for producing zeolite separation membrane. U.S. Patent 8,263,179 2012.
  • 261.Kazemimoghadam M, Mohammadi T. Preparation of nano pore hydroxysodalite zeolite membranes using of kaolin clay and chemical sources. Desalination. 2011;278(1–3):438–42. [Google Scholar]
  • 262.Khajavi S, Jansen JC, Kapteijn F. Performance of hydroxy sodalite membranes as absolute water selective materials under acidic and basic conditions. J Membr Sci. 2010;356:1–6. [Google Scholar]
  • 263.Satou K, inventor; Mitsubishi Chemical Corporation, assignee. Zeolite membrane, separation membrane, and component separation method. U.S. Patent 8,721,891 2014.
  • 264.Goh PS, Ismail AF, Sanip SM, Ng BC, Aziz M. Recent advances of inorganic fillers in mixed matrix membrane for gas separation. Sep Purif Technol. 2011;81(3):243–64. [Google Scholar]
  • 265.Gao Z, Yue Y, Li W. Application of zeolite-filled pervaporation membrane. Zeolites. 1996;16(1):70–4. [Google Scholar]
  • 266.Crank J The mathematics of diffusion. Bristol, UK: J.W. Arrowsmith Ltd.; 1975. [Google Scholar]
  • 267.Petropoulos JH. A comparative study of approaches applied to the permeability of binary composite polymeric materials. Journal of Polymer Science: Polymer Physics Edition. 1985;23(7):1309–24. [Google Scholar]
  • 268.Malekpour A, Mostajeran B, Koohmareh GA. Pervaporation dehydration of binary and ternary mixtures of acetone, isopropanol and water using polyvinyl alcohol/zeolite membranes. Chem Eng Process: Process Intensification. 2017;118:47–53. [Google Scholar]
  • 269.Huang Z, Guan HM, Tan WL, Qiao XY, Kulprathipanja S. Pervaporation study of aqueous ethanol solution through zeolite-incorporated multilayer poly(vinyl alcohol) membranes: Effect of zeolites. J Membr Sci. 2006;276:260–71. [Google Scholar]
  • 270.Khoonsap S, Amnuaypanich S. Mixed matrix membranes prepared from poly (vinyl alcohol) (PVA) incorporated with zeolite 4A-graft-poly (2-hydroxyethyl methacrylate) (zeolite-g-PHEMA) for the pervaporation dehydration of water- acetone mixtures. J Membr Sci. 2011;367:182–9. [Google Scholar]
  • 271.Oh D, Lee S, Lee Y. Mixed-matrix membrane prepared from crosslinked PVA with NaA zeolite for pervaporative separation of water–butanol mixtures. Desalination and Water Treatment. 2013;51(25–27):5362–70. [Google Scholar]
  • 272.Huang Z, Shi Y, Wen R, Guo Y-H, Su J-F, Matsuura T. Multilayer poly(vinyl alcohol)–zeolite 4A composite membranes for ethanol dehydration by means of pervaporation. Sep Purif Technol. 2006;51(2):126–36. [Google Scholar]
  • 273.Amirilargani M, Sadatnia B. Poly(vinyl alcohol)/zeolitic imidazolate frameworks (ZIF-8) mixed matrix membranes for pervaporation dehydration of isopropanol. J Membr Sci. 2014;469(0):1–10. [Google Scholar]
  • 274.Wu G, Jiang M, Zhang T, Jia Z. Tunable Pervaporation Performance of Modified MIL-53(Al)-NH2/Poly(vinyl Alcohol) Mixed Matrix Membranes. J Membr Sci. 2016;507:72–80. [Google Scholar]
  • 275.Hung W-S, Chang S-M, Lecaros RLG, Ji Y-L, An Q-F, Hu C-C, et al. Fabrication of hydrothermally reduced graphene oxide/chitosan composite membranes with a lamellar structure on methanol dehydration. Carbon. 2017;117:112–9. [Google Scholar]
  • 276.Lecaros RLG, Mendoza GEJ, Hung W-S, An Q-F, Caparanga AR, Tsai H-A, et al. Tunable interlayer spacing of composite graphene oxide-framework membrane for acetic acid dehydration. Carbon. 2017;123:660–7. [Google Scholar]
  • 277.Zheng S, Wang P, Romero R, Kitahara I, Lin W, inventors; Nitto Denko Corporation, assignee. Selectively permeable graphene oxide/polyvinyl alcohol membrane for dehydration WO Patent Application WO2017044845A1 2017.
  • 278.Kujawski J, Rozicka A, Bryjak M, Kujawski W. Pervaporative removal of acetone, butanol and ethanol from binary and multicomponent aqueous mixtures. Sep Purif Technol. 2014;132:422–9. [Google Scholar]
  • 279.Sato K, Sugimoto K, Nakane T. Preparation of higher flux NaA zeolite membrane on asymmetric porous support and permeation behavior at higher temperatures up to 145 °C in vapor permeation. J Membr Sci. 2008;307(2):181–95. [Google Scholar]
  • 280.Ortiz I, Gorri D, Casado C, Urtiaga A. Modelling of the pervaporative flux through hydrophilic membranes. J Chem Technol Biotechnol. 2005;80(4):397–405. [Google Scholar]
  • 281.van Hoof V, Dotremont C, Buekenhoudt A. Performance of Mitsui NaA type zeolite membranes for the dehydration of organic solvents in comparison with commercial polymeric pervaporation membranes. Sep Purif Technol. 2006;48:304–9. [Google Scholar]
  • 282.Gallego-Lizon T, Edwards E, Lobiundo G, dos Santos LF. Dehydration of water/t -butanol mixtures by pervaporation: comparative study of commercially available polymeric, microporous silica and zeolite membranes. J Membr Sci. 2002;197(1–2):309–19. [Google Scholar]
  • 283.Ong YK, Shi GM, Le NL, Tang YP, Zuo J, Nunes SP, et al. Recent Membrane Development for Pervaporation Processes. Progress in Polymer Science. 2016;57:1–31. [Google Scholar]
  • 284.Nebraska-based E Energy Adams to install Whitefox ICE 2018. [Available from: http://www.ethanolproducer.com/articles/15262/].
  • 285.Portz T Outfoxing dehydration Ethanol Producer Magazine. 2017. [Google Scholar]
  • 286.Albrecht T ‘The Way Forward’ Ethanol Producer Magazine. 2018:18–25. [Google Scholar]
  • 287.Voegele E Aemetis to add Zebrex membrane technology at Keyes plant Ethanol Producer Magazine 2018. [Available from: http://ethanolproducer.com/articles/15587/aemetis-to-add-zebrex-membrane-technology-at-keyes-plant].
  • 288.Smallwood IM. Handbook of organic solvent properties. New York: Halsted Press; 1996. [Google Scholar]
  • 289.Nishikawa M, Kotoh K, Koga K, Katase A. Fractional Treatment of Spent Scintillation Cocktail with Low Level Tritium by Azeotropic Distillation. Journal of Nuclear Science and Technology. 1982;19(11):928–35. [Google Scholar]
  • 290.Hazardous Air Pollutants: U.S. Environmental Protection Agency; 2018. [Available from: https://www.epa.gov/haps].
  • 291.Yaws CL. Yaws’ Critical Property Data for Chemical Engineers and Chemists. Online version available at: https://app.knovel.com/hotlink/toc/id:kpYCPDCECD/yaws-critical-property/yaws-critical-property. Knovel; 2012: 2013: 2014. [Google Scholar]
  • 292.Senichev VY, Tereshatov VV. Chapter 4. General Principles Governing Dissolution of Materials in Solvents. Section 4.1 Simple solvent characteristics In: Wypych G, editor. Handbook of Solvents: ChemTec Publishing; 2001. [Google Scholar]
  • 293.Knovel Solvents - A Properties Database. ChemTec Publishing; 2012. [Internet]. Available from: https://app.knovel.com/hotlink/toc/id:kpKSAPD005/knovel-solvents-properties/knovel-solvents-properties. [Google Scholar]
  • 294.Thiess H, Schmidt A, Strube J. Development of a Scale-up Tool for Pervaporation Processes. Membranes. 2018;8(1):4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 295.Van Baelen D, Reyniers A, Van der Bruggen B, Vandecasteele C, Degreve J. Pervaporation of binary and ternary mixtures of water with methanol and/or ethanol. Sep Sci Technol. 2004;39:563–80. [Google Scholar]
  • 296.Haaz E, Jozsef Toth A. Methanol dehydration with pervaporation: experiments and modelling. Sep Purif Technol. 2018;205:121–9. [Google Scholar]
  • 297.Kita H, Sasaki S, Tanaka K, Okamoto K-i, Yamamoto M. Esterification of Carboxylic Acid with Ethanol Accompanied by Pervaporation. Chemistry Letters. 1988;17(12):2025–8. [Google Scholar]
  • 298.PERVATECH BV Datasheet: Zeolite NaA Membranes 2017.
  • 299.Kreiter R, Rietkerk MDA, Castricum HL, van Veen HM, ten Elshof JE, Vente JF. Evaluation of hybrid silica sols for stable microporous membranes using high-throughput screening. Journal of Sol-Gel Science and Technology. 2011;57(3):245–52. [Google Scholar]
  • 300.Chang K-S, Yoshioka T, Kanezashi M, Tsuru T, Tung K-L. A molecular dynamics simulation of a homogeneous organic-inorganic hybrid silica membrane. Chemical Communications. 2010;46(48):9140–2. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplement1

RESOURCES