Elsevier

Journal of Membrane Science

Volume 287, Issue 2, 15 January 2007, Pages 162-179
Journal of Membrane Science

Review
Polymeric membrane pervaporation

https://doi.org/10.1016/j.memsci.2006.10.043Get rights and content

Abstract

Pervaporation is an efficient membrane process for liquid separation. The past decades had witnessed substantial progress and exciting breakthroughs in both the fundamental and application aspect of pervaporation. This review provided an analytical overview on the potential of pervaporation for separating liquid mixtures in terms of the solubility parameter and the kinetic parameter of solvents. Focus of the review was given to the fundamental understanding of the membrane. Research progress, challenges and opportunities, and the prospect of pervaporation were also discussed. The thermodynamic approach of pervaporation, featuring emphasizing membrane/species interactions, though gained great successes in the past decades, is now facing its toughest challenge in the org–org separation. A kinetic era of pervaporation, featuring emphasizing diffusion selectivity, as well as the synergy between the selective diffusion and sorption, is in the making, and this approach will eventually find solutions to the challenging org–org separation.

Introduction

Pervaporation is a membrane process for liquid separation [1], [2], a polymeric or zeolite membrane [3], [4], [5] usually serves the separating barrier for the process. When a membrane is in contact with a liquid mixture, one of the components can be preferentially removed from the mixture due to its higher affinity with, and/or quicker diffusivity in the membrane. As a result, both the more permeable species in the permeate, and the less permeable species in the feed, can be concentrated. In order to ensure the continuous mass transport, very low absolute pressures (e.g., 133.3–400.0 Pa (1–3 mmHg)) are usually maintained at the downstream side of the membrane, removing all the molecules migrating to the face, and thus rendering a concentration difference across the membrane. As a variant, the use of a sweeping gas [6], [7] in the downstream side of the membrane is also a feasible alternative for the generally used vacuum operation.

It is well known that the phase change from liquid to vapor takes place in pervaporation. Processes involving phase changes are generally energy-intensive, and distillation is a notorious example of them. Pervaporation cleverly survives the challenge of phase change by two features. (1) Pervaporation deals only with the minor components (usually less than 10 wt.%) of the liquid mixtures, and (2) pervaporation uses the most selective membranes. The first feature effectively reduces the energy consumption of the pervaporation process. Compared with the distillation, because of the characteristics of pervaporation operation, it is essentially true that only the minor component in the feed consumes the latent heat. The second feature generally allows pervaporation the most efficient liquid-separating technology. Take the separation of isopropanol/water mixtures for example, if the water content in the feed is 10 wt.%, the maximum single plate separation factor (isopropanol to water) in distillation is about 2, however, a pervaporation membrane can normally offer an one-through separation factor (water over isopropanol) of 2000–10,000 [8], [9], [10]. Furthermore, combination of these two features ranks pervaporation the most cost-effective liquid separation technology [11], [12]. In addition, pervaporation also demonstrates incomparable advantages in separating heat-sensitive, close-boiling, and azeotropic mixtures [13], [14], [15], [16] due to its mild operating conditions, no emission to the environment, and no involvement of additional species into the feed stream. More recently, the hybrid processes [12], [17], [18], [19] integrating pervaporation with other viable liquid-separating technologies, and processes are gaining momentum. With these developments, we have more reasons to believe that pervaporation will play even more important roles in the future.

To date, pervaporation has found viable applications [20] in the following three areas: (i) dehydration of organic solvents (e.g., alcohols, ethers, esters, acids); (ii) removal of dilute organic compounds from aqueous streams (e.g., removal of volatile organic compounds, recovery of aroma, and biofuels from fermentation broth); (iii) organic–organic mixtures separation (e.g., methyl tert-butyl ether (MTBE)/methanol, dimethyl carbonate (DMC)/methanol). Among them, dehydration of organic solvents is best developed. This resulted from the so-called synergic effect [21]: water is both preferentially dissolved and transported in the hydrophilic membranes due to its much smaller molecular size. When pervaporation is used for removing organic compounds from water, the preferential transport of the organic species cannot be achieved in the organophilic membrane. As a result, the permselectivity of the pervaporation process is reduced, and less than the sorption selectivity. Theoretically speaking, pervaporation in these cases demonstrates no advantage over the adsorption technique. However, when the concentration of organic compounds in water is relatively high, pervaporation tends to be superior to the adsorption technology since pervaporation is a continuous process, it therefore suffers no limitation of the saturated adsorption capacity, which is however an intrinsic weakness of the adsorption process.

Separation of organic–organic mixtures represents the most challenging application for pervaporation [22]. Most liquid pairs in this category are of industrial importance [19], such as polar/non-polar, e.g., methanol/MTBE [23], [24], aromatic/aliphatic, e.g., benzene/n-hexane [25], [26], aromatic/alicyclic, e.g., benzene/cyclohexane [27], [28], and isomers, e.g., p-xylene, m-xylene, and o-xylene [29], [30], [31]. Research in this category gained some successes in the separation of polar/non-polar liquid mixtures as shown previously, but has not yet seen much significant progress in the other liquid pairs.

To date, several reviews [2], [11], [12], [17], [20], [22], [32], [33] on pervaporation have been available. A detailed review on the zeolite membrane pervaporation had also been conducted by Bowen et al. [11], this review will thus focus on the polymeric membrane pervaporation, with the emphasis given to the fundamental understanding of the membranes, where an analytical overview on the potential of the pervaporation for separating various liquid pairs was presented, the challenges and opportunities, and the prospect of pervaporation was also tentatively discussed.

Section snippets

Pervaporation in terms of the Hansen solubility parameter and the kinetic diameter

If reviewing the evolution of pervaporation, one may recognize the fact that the Hansen solubility parameter exerted incomparable influence on the development of pervaporation. The Hansen solubility parameter refers to the density of cohesive energy [34], which consists of three components: δh: the contribution of the hydrogen bonding interaction; δp: the contribution of the polar interaction; δd: the contribution of the dispersion interaction. The solubility parameter can thus be represented

Solution-Diffusion theory

Solution-diffusion is the generally accepted mechanism of mass transport through non-porous membranes, which was first proposed by Graham [38] based on his extensive research on gas permeation through homogeneous membranes. It is held that gas permeation through a homogeneous membrane consists of three fundamental processes: (1) solution of gas molecules in the upstream surface of the membrane. (2) Diffusion of the dissolved species across the membrane matrix. (3) Desorption of the dissolved

A closer view of the pervaporation membranes

Generally speaking, because of the presence of strong membrane–species interactions, the pervaporation membrane can no longer be treated as a uniform medium for permeation. Shimidzu and co-workers [62], [63], [64] believed that the polar groups in the membrane matrix, responsible for the membrane hydrophilicity, act as the fixed carriers for mass transport in the membrane. In the case of dehydration of organic solvents, it is believed that water transport in the membrane proceeds in a special

The coupled transport in pervaporation membranes

The coupled transport is a frequently observed mass transport phenomenon in pervaporation membranes [68], [69], [70], [71]. According to Mulder et al. [72], there are generally two types of coupled transport, i.e., the thermodynamic, and kinetic coupling. The thermodynamic coupling results from the interaction between the dissolved species in the membrane. As is well known that the Gibbs free energy of one species can be changed by the presence of other species, and the changes in the free

Structural stability of composite pervaporation membranes

Pervaporation membranes fall into two categories, the homogeneous membrane and the composite membrane. The composite membrane can offer a higher permeation flux than the homogeneous one due to the much thinner thickness of the homogeneous membrane supported on a porous substrate. The composite membrane is thus suitable for industrial use. Ideally, the porous substrate of a composite membrane presents negligible resistance to mass transport [73]. Otherwise, the substrate resistance leads to

Organic solvent dehydration

Dehydration of organic solvents (e.g., alcohols, ethers, acids, and ketones) largely represents the applications of the pervaporation [90], [91], [92], [93], [94]. The materials used in earlier dehydration research were the naturally occurring polymers, e.g., cellulose and cellulose derivatives [1]. Synthetic polymers [95], [96], e.g., poly(acrylic acid) (PAA), poly(vinyl alcohol) (PVA), polyacrylonitrile (PAN), and nylon 6, were subsequently investigated, with a focus mainly on ethanol

Commercial and engineering aspects of pervaporation

Since the commercialization of pervaporation for ethanol dehydration launched by GFT in 1980s based on the cross-linked PVA/PAN composite membrane, both the scope of application and the types of the pervaporation membranes were extensively enlarged [166]. According to the website of Sulzer Chemtech, a wide array of solvents have been covered in its dehydration market, which includes various alcohols, ethers, ketones, acids, and some polymer solvents like THF, dioxane, etc. The SULTZER PERVAP®

Concluding remarks

Pervaporation has played an important role in solvent dehydration, and this application can be further extended by integrating pervaporation with other viable liquid-separating technologies, and by finding right materials for dehydrating some caustic solvents (e.g., nitric acid). Higher membrane productivity and selectivity is always a concern for application, and this was conventionally accomplished by operating the pervaporation membranes at higher temperatures. The improved membrane flux can

Acknowledgement

The financial support from Natural Sciences and Engineering Research Council (NSERC) of Canada is gratefully acknowledged.

References (168)

  • K. Inui et al.

    Permeation and separation of benzene cyclohexane mixtures through benzoylchitosan membranes

    J. Membr. Sci.

    (1998)
  • H.L. Chen et al.

    PVA membrane filled beta-cyclodextrin for separation of isomeric xylenes by pervaporation

    Chem. Eng. J.

    (2000)
  • M. Schleiffelder et al.

    Crosslinkable copolyimides for the membrane-based separation of p-/o-xylene mixtures

    React. Funct. Polym.

    (2001)
  • K. Wegner et al.

    Polycrystalline MFI zeolite membranes: xylene pervaporation and its implication on membrane microstructure

    J. Membr. Sci.

    (1999)
  • S.I. Semenova et al.

    Hydrophilic membranes for pervaporation: an analytical review

    Desalintion

    (1997)
  • M. Mulder et al.

    Preferential sorption versus preferential permeability in pervaporation

    J. Membr. Sci.

    (1985)
  • H.K. Lonsdale

    The growth of membrane and technology

    J. Membr. Sci.

    (1982)
  • W. Ji et al.

    Modeling of multicomponent pervaporation for removal of volatile organic compounds from water

    J. Membr. Sci.

    (1994)
  • I. Blume et al.

    Separation of dissolved organics from water by pervaporation

    J. Membr. Sci.

    (1990)
  • M.R. Coleman et al.

    Isomeric polyimides based on fluorinated dianhydrides and diamines for gas separation applications

    J. Membr. Sci.

    (1990)
  • J.P. Brun et al.

    Modeling of the pervaporation of binary mixtures through moderate swelling, non-reacting membranes

    J. Membr. Sci.

    (1985)
  • T. Okada et al.

    A new transport model for pervaporation

    J. Membr. Sci.

    (1991)
  • T. Okada et al.

    A study on pervaporation of ethanol/water mixtures on the basis of pore flow model

    J. Membr. Sci.

    (1991)
  • M. Yoshikawa et al.

    Polymer membrane as a reaction field, II effect of membrane polarity on selective separation of water-ethanol binary mixtures through synthetic polymer membranes

    J. Membr. Sci.

    (1986)
  • I. Cabasso et al.

    The permselectivity of ion-exchange membranes for non-electrolyte liquid mixtures, I. Separation of alcohol/water mixtures with Nafion hollow fibers

    J. Membr. Sci.

    (1985)
  • R.Y.M. Huang et al.

    Pervaporation separation of water/isopropanol mixture using the sulfonated poly(ether ether ketone) (SPEEK) membranes: transport mechanism and separation performance

    J. Membr. Sci.

    (2001)
  • O. Kedem

    The role of coupling in pervaporation

    J. Membr. Sci.

    (1989)
  • M. She et al.

    Effects of concentration, temperature, and coupling on pervaporation of dilute flavor organics

    J. Membr. Sci.

    (2006)
  • S. Tan et al.

    Pervaporation of alcoholic beverages—the coupling effects between ethanol and aroma compounds

    J. Membr. Sci.

    (2005)
  • M.H.V. Mulder et al.

    Ethanol-water separation by pervaporation

    J. Membr. Sci.

    (1983)
  • G.H. Koops et al.

    Poly(vinyl chloride) polyacrylonirile composite membranes for the dehydration of acetic acid

    J. Membr. Sci.

    (1993)
  • W. Xu et al.

    Carboxylic acid containing polyimides for pervaporation separation of toluene/iso-octane mixtures

    J. Membr. Sci.

    (2003)
  • S. Matsui et al.

    Pervaporation of aromatic/aliphatic hydrocarbons by crosslinked poly(methyl acrylate-acrylic acid) membranes

    J. Membr. Sci.

    (2002)
  • S. Matsui et al.

    Pervaporation separation of aromatic/aliphatic hydrocarbons by crosslinked poly(n-alkyl acrylate) membranes

    J. Membr. Sci.

    (2003)
  • R.Y.M. Huang et al.

    Characteristics of sodium alginate membranes for the pervaporation dehydration of ethanol-water and isopropanol-water mixtures

    J. Membr. Sci.

    (1999)
  • R.Y.M. Huang et al.

    Pervaporation separation of aqueous mixtures using crosslinked poly(vinyl alcohol). 2. Permeation of ethanol-water mixtures

    J. Membr. Sci.

    (1990)
  • C.K. Yeom et al.

    Vapor permeation of ethanol-water mixtures using sodium alginate membrane with crosslinking gradient structure

    J. Membr. Sci.

    (1997)
  • R.Y.M. Huang et al.

    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)
  • R.Y.M. Huang et al.

    Resistance model approach to asymmetric polyetherimide membranes for pervaporation of isopropanol/water mixtures

    J. Membr. Sci.

    (1993)
  • P. Shao et al.

    Composite membranes with an integrated skin layer: preparation, structural characteristics and pervaporation performance

    J. Membr. Sci.

    (2005)
  • Y. Zhu et al.

    A continuous pervaporation membrane reactor for the study of esterification reactions using composite polymeric/ceramic membranes

    Chem. Eng. Sci.

    (1996)
  • U. Sander et al.

    Design and operation of a pervaporation plant for ethanol dehydration

    J. Membr. Sci.

    (1988)
  • P. Aptel et al.

    Applications of pervaporation to separate azeotropic mixtures

    J. Membr. Sci.

    (1976)
  • M. Kondo et al.

    Tubular-type pervaporation module with zeolite NaY membranes

    J. Membr. Sci.

    (1997)
  • J. Caro et al.

    Zeolite membranes-state of their development and perspective

    Micropor. Mesopor. Mater.

    (2000)
  • Y. Morigami et al.

    The first large-scale pervaporation plant using the tubular-type module with zeolite NaA membrane

    Sep. Purif. Technol.

    (2001)
  • J. Neel
  • X. Feng et al.

    Liquid separation by membrane pervaporation: a review

    Ind. Eng. Chem. Res.

    (1997)
  • J. Neel

    Pervaporation: fundamentals and practice

    Makromol. Chem. Macromol. Symp.

    (1993)
  • G.H. Koops et al.
  • Cited by (0)

    View full text