You are on page 1of 14

No dictionary results were found. Please try another search.

Pervaporation
December 4, 1998 ENVE 436 Sonja Thoma Andrea Resch Lanaya Voelz

TABLE OF CONTENTS

History Membranes Principles System Design Application Case Study Conclusion Bibliography

1 1 2 3 4 5 6 7

Appendix Figure 1: Permeate pressure vs. water flux Figure 2: Once through pervaporation system Figure 3: Batch pervaporation system Figure 4: Schematic diagram of the automated pervaporation unit Table 1: Comparative Pervaporation Performances of Various Membrane Materials (purchased) Table 2: Comparative Pervaporation Performances of Various Membrane Materials (prepared) 8 8 9 9 10 10

HISTORY As early as 1906, L. Kahlenberg reported a qualitative study on the separation of a mixture of a hydrocarbon and an alcohol through a rubber membrane. In 1955, D. H. Hagerbaumer conducted the first quantitative investigation with a microporous Vycor glass membrane with a high-pressure drop across it, to allow for the separation of liquid-liquid mixtures.(1,99) Later in 1965 Binning et al., at American Oil, utilized this operation of separating a liquid-liquid mixture into a vapor mixture using a nonporous polymeric film. This research yielded a high degree of separation along with high permeation rates. The process did not come into commercial use until 1982 when Gesellchaft fur Trenntechnik mbH (GFT) of Germany installed a pervaporation plant to separate water from concentrated alcohol solutions. Since then, more than one hundred plants have been installed. Recently, Exxon has used pervaporation in its refineries to separate hydrocarbon mixtures containing aromatics and aliphatics. Another commercial use for pervaporation is the removal of methylene chloride from small waste streams.(2,34) MEMBRANE CHARACTERISTICS Synthetic membranes are thin, solid-phase barriers that allow preferential passage of certain substances under the influence of a driving force. Both the chemical and the physical nature of the membrane material control membrane separation. Membrane separation occurs because of differences in size, shape, chemical properties, or electrical charge of the substances to be separated. Microporous membranes control separation by size, shape and charge discrimination, whereas nonporous membranes

depend on sorption and diffusion. The performance of the membrane is determined by the degree of separation of fluid mixtures and permeation rate (flux).(3,33) Three general categories of inorganic membranes are ceramics, metals and glass. Because they are so rigid, ceramic microfilters accommodate fluxes five to ten times greater than those of asymmetric polymeric membranes. They can be backwashed frequently without damaging the membrane skin layer. Ceramic membranes are highly resistant to cleaning chemicals and can be sterilized repeatedly by high pressurized steam. Their life span is up to ten years compared to the typical life spans for polymer membranes, which are about one year for hydrophobic membranes and up to four years for fluoropolymers. Ceramic membranes are brittle and are more expensive than polymeric membranes.(3,34) Pervaporation membranes are typically composites. The first layer is a porous, polymeric support coated with a second polymer, the "active" or "permselective" layer, which is engineered to preferentially absorb the chemical species of interest. The membranes separation characteristics can be further refined by varying the thickness of the permselective layer.(4,88) For example, asymmetric composite hydrophilic membranes such as composite PVA-PS (Poly(vinyl alcohol)-Polysulfone) are used for pervaporation. Pervaporation separation plants contain between ten and one hundred m2 of membrane area, which must be packaged efficiently and economically into units called membrane modules. Flat-sheet and spiral-wound modules are commonly used.(3,41) Silicon rubber membranes are also used in pervaporation. Spiral wound configuration offers a high membrane surface area per module and allows for relatively high feed flow rates which is common for pervaporation. Silicone rubber pervaporation modules are remarkably effective at separating organic solutes from dilute aqueous solutions.(2,35) Spiral wound modules are available in 2, 4 and 8 inch diameters to accommodate a variety of feed flow requirements and to allow for economical system design.(4,89) The choice membrane depends on the feed solution. The most efficient application of any membrane is to permeate the minor component of a mixture.(2,35) PERVAPORATION PRINCIPLES Pervaporation is based on selective separation of a feed liquid. "Pervaporation is unique among membrane separation techniques, in that a phase change takes place across the membrane.." (4,88) Removal of a certain compound is accomplished by a the partial pressure difference created on feed and permeate sides of the membrane. The separation is a function of the rate of permeation of the components of the mixture through the membrane. Different types of polymer membranes allow for the adsorption of different chemical species. One membrane may be designed to remove

water from the feed stream, while another may be designed to remove a certain hydrocarbon. The principles of pervaporation can be best understood through explanation of a two step process, an evaporation process and a membrane transfer process. In the evaporation process, the temperature of the feed liquid is elevated to the point where a saturated vapor is formed. When the feed enters the apparatus, the saturated vapor comes in contact with the membrane. Mathematically speaking, this process is "defined as the ratio of the component concentration in the feed vapor to their concentration in the feed liquid."(2,35) In the second process, the vapor diffuses from the feed side, across the membrane, to the permeate side. A condenser is installed on the permeate side to create a area of pressure lower than the feed side. Mathematically speaking, the second process can be defined as "the ratio of the components in the permeate vapor to the ratio of the components in the feed vapor."(2,36) The mass flux is due to "the continuous adsorption on one side (of the membrane) and desorption on the other."(4,88) Mass flux across the membrane can be described as a partial pressure difference between the feed saturated vapor pressure and the permeate vapor pressure. Experimental data for water flux across a 20- m-thick silicone rubber pervaporation membrane shows that as the permeate pressure decreases and approaches the feed pressure, the flux decreases to zero in a linear matter. (See figure 1) Temperature is also a factor in this flux. As the temperature of the feed is increased, the system is more able to handle a higher flux of water. This is the reason most feed streams pass though a heater before entering the membrane housing. The final separation of the feed liquid is the product of the separation achieved by the evaporation of the liquid and the separation achieved by the permeation through the membrane. Since the separation factor of a contaminated feed stream is based upon the product of the evaporation factor and the membrane factor, the separation factor may be due more to one factor than another. For example, "the 200- to 500-fold separation achieved by pervaporation membranes in ethanol dehydration is entirely attributed to the separation factor of the membrane which is more permeable to water than ethanol." (2.37) This means that the membrane factor is greater than the evaporation factor. On the other hand, when attempting to separate VOCs from a dilute aqueous solution, the evaporation factor is greater than the membrane factor. The separation factor for a VOC will increase the more hydrophobic it becomes, but the contribution to the total separation due to membrane will decrease. SYSTEM DESIGN

The plant design for a pervaporation system depends of the particular stream being treated. For example, the once through system is best applied to solutions containing water-soluble VOCs with separation factors between 10 and 500, for which concentration polarization is not a major problem.(2,37) The separation factor or system selectivity, is the ratio of mass fraction of a permeating component on the permeate side of the membrane, compared to the feed side. = [HC]P/[H2O]P [HC]F/[H2O]F where HC is the concentration (in grams or moles) of hydrocarbons in the permeate (P) or feed (F). The extent to which a system is effective is indicated by its separation factor. The less volatile and more soluble a chemical is in water, the lower the separation factor is. For example, alcohols and aldehydes have low separation factors. High separation factor chemicals, such as aromatics and chlorinated compounds, have higher volatility and are less soluble in water. The higher the separation factor is for a chemical, the easier it is to separate out of solution with pervaporation.(4,88) In once through systems, the membrane modules are arranged in series, and the feed solution is discharged after it passes through the modules. (See figure 2) The number of modules in series and the feed flow rate determines the velocity of the solution. For VOCs with modest separation factors, such a system can control concentration polarization and sufficient residence time within the module to remove the required VOC from the feed. For VOCs with large separation factors, such as toluene and tricholoroethylene, matching the fluid velocity required to control concentration polarization with the residence time required to achieve target VOC removal in a single pass becomes more difficult. Many modules may be needed, requiring booster pumps to push the feed solution through the modules.(2,37-38) For treating water containing VOCs with separation factors greater than 500, the batch system design may be more suitable. (See figure 3) In a batch system, feed solution is accumulated in a surge tank. Some of this solution is then transferred to the feed tank and circulated at high velocity through the pervaporation modules until the VOC concentration reaches the desired level. The treated water is removed from the feed tank, and the tank is loaded with a new batch of untreated solution and the cycle is repeated.(2,38) Pervaporation system operation is completely automatic and controlled by a programmable logic controller. The cycle time of the controller is adjustable and can

be set to achieve any desired degree of VOC removal or the same degree of removal with different VOCs. At higher temperatures, pervaporation systems have a higher performance, but that needs to be balanced against the cost of the energy required to heat the feed.(2,38) APPLICATIONS Extraction by partial vaporization through a dense membrane, pervaporation, has many advantages, including:

elimination of membrane stability problems; preclusion of fouling by microorganisms; absence of thermal, chemical or mechanical stress on a fermentation broth; increase in productivity through increased glucose consumption and alcohol production (5, 222).

An example of pervaporation in treatment of wastewater is the removal of methyl ethyl ketone (MEK) and benzene. When comparing the overall economics of recovery for steam stripping, chemical oxidation, and pervaporation, pervaporation was found to be the least expensive. Pervaporation was shown to enrich the organic concentration in the plants waste stream by a factor of over 100. The residual water stream contained 2 ppm benzene, which complied with permit levels allowing discharge into a local sewer (4, 89). Another case describes the removal of trichloroethylene (TCE) from ground water at a former metal parts manufacturing site via pervaporation. In this case, pervaporation reduced TCE volumes from 100 ppm to less than 0.1 ppm (4, 89). Other major applications of pervaporation include the removal of small amounts of water from organic solutions like the drying of isopropanol-water, butanol water or ethanol-water azeotropes to produce a relatively pure organic chemical. Ethanol purity levels of 99.85% are attained in ethanol-water separation with a distillation pervaporation hybrid process. Pervaporation is also used in the drying of organic liquids by removing small volumes of water from nearly pure organics (3, 41). Although pervaporation demonstrates many advantages as an industrial membrane process, including economical advantages, overall, commercial sales of pervaporation systems are low because of the high cost of reheating the permeate between stages to sustain a meaningful flux and because of the large membrane surface area required (3, 41) .

CASE STUDY In the Journal of Chemical Technology and Biotechnology, Eric Favre, Quang Trong Nguyen and Sophia Bruneau published results and conclusions from extraction of 1Butanol from aqueous solutions by pervaporation. The predominance of driving force on pervaporation extraction performances was shown by a comparative study on different binary aqueous solutions of alcohols (5, 221). The production of solvents from biological waste fermentation suffers from the low productivity of bioprocesses, particularly for the acetone-butanol-ethanol (ABE) fermentation process. Since it cannot exceed 1% (v/v) final solvent concentration in broth, 1-butanol produces an inhibitory effect on Clostridium acetobutylicum. This effect may be overcome by extractive fermentation of 1-butanol and recently, dense membrane processes, pervaporation, is one of the various integrated extraction strategies that have been studied (5, 221). The removal of 1-butanol from aqueous streams can be achieved providing a hydrophobic and highly permeable membrane material is used. Prior to this study, few membrane materials have been tested for this particular application. The objectives of this work include the following:

comparison of the performances of membrane materials screening of polymetric materials to identify improved membrane candidates to better understand the influence of system variable on pervaporation performance

For the experiments reported in this work, simple binary and ternary mixtures were used to enable more clear interpretation of the pervaporation transport phenomenon (5, 222) . Dense membranes were either purchased from a supplier, or prepared in the laboratory. The results from the comparative pervaporation performances of various membrane materials tested for 1-Butanol extraction from water are shown in Table 1 (purchased from supplier) and Table 2 (prepared in the laboratory). Figure 4 is a schematic diagram of the automated pervaporation unit used for flux and selectivity determinations. The pervaporation experiments were performed at 40 C in a flat module with a membrane surface area of 4 cm2. The vacuum was maintained by a pump below 200 Pa at the downstream side of the membrane. Total volume of the feed mixture was 300 cm3 and consisted of 1 % (w/w) 1-butanol aqueous solution. The feed mixture was recirculated through a temperature bath at a high flow rate to prevent concentration polarization in the module. The vapors transferred through the

membrane were automatically collected in a cold trap, thawed and analyzed in a gas chromatograph to determine alcohol and water content (5, 223). The polydimethylsiloxane (PDMS) membranes, which have been used for the extraction of hydrocarbons, halogenated hydrocarbons, esters and ketones from water, prove superior in terms of the compromise between flux and selectivity. Although PDMS was identified as being the best dense membrane material for 1-butanol extraction, its performance may be improved by adding non-polymetric compounds. Further investigation found that absorbent-filled PDMS membranes demonstrated significant improvements in terms of selectivity, but decreased permeability. Oleyl alcohol-impregnated PDMS membranes demonstrated pervaporation performances greatly improved initially, however this improvement was not maintained. This situation may have resulted from either oleyl alcohol vaporization in the vapor phase, or release into the liquid phase side or minute amounts of 1-butanol that could significantly increase the solubility of oleyl alcohol in water (5, 224). The analysis of the membrane transport properties of a pervaporation process is based on the "solution-diffusion" mechanism. The sorption equilibrium at the upstream part of the membrane (S) and the diffusion of permeates in the polymer network (D) make up the two steps of the solution diffusion mechanism. These two steps contribute to permeability (P) and pervaporation flux that can be interpreted as the result of both permeability and driving force such as concentration gradient. J = (D x S x c)/z = (P x c)/z
(5, 223)

Much of the conclusions of this study are hypothetical. If the qualitative relationship between pervaporation fluxes and feed activities can be shown to apply on a general basis for dilute solutions through PDMS membranes, it could facilitate the prediction of the influence of one component of the feed mixture on extraction performances (5, 224) . CONCLUSION Pervaporation is new technology - the first commercial plants for dehydration of alcohol were installed only 10 years ago. Recently, the separation of VOCs from aqueous waste streams, the removal of small amounts of water from organic solutions and the drying of organic liquids by removing small volumes of water appear to be promising applications. The process is most applicable to recovery of VOCs that are more hydrophobic than acetone in streams that contain VOC levels between 200 and 50,000 ppm and have a volumetric flow of 10 to 100 gal per minute. In this range,

pervaporation offers significant economic and technical advantages over alternative processes.(2,38-39)

BIBLIOGRAPHY 1. Hwang, S. and K. Kammermyer. Membranes in Separation. John Wiley and Sons, Inc. 1975, 99-123. 2. Athayde, A.L., R. W. Baker, R. Daniels, M. H. Le and J.H. Ly. Pervaporation for Wastewater Treatment. Chemtech. January, 1997, 34-39. 3. Singh, Rajindar. Industrial Membrane Separation Processes. Chemtech. April, 1998, 33-44. 4. Barber, Terry and Brian D. Miller. Pervaporation Technology: Fundamentals and Environmental Applications. Chemical Engineering. September, 1994, 8890. 5. Favre, Eric, Quang Trong Nguyen and Sophie Bruneau. Extraction of 1Butanol from Aqueous Solutions by Pervaporation. Journal of Chemical Technology and Biotechnology. 1996, 65, 221-228.

Appendix Figure 1: Permeate pressure vs. water flux

Permeate pressure, cmHg

Figure 2: Once through pervaporation system

Figure 3: Batch pervaporation system

Figure 4: Schematic diagram of the automated pervaporation unit

You might also like