TRANSPORT IN MEMBRANES 189 feed permeate phase 1 membrane phase 2 m 11I, ,s1 ms P2 11mI, , 1 s 1li,2 ~,2 PI Ci,l Ci,l ms Ci,2 ci,2 sm ami,2 8 a i,2 ~,1 a'I1, Figure V - 23. Process conditions for transport through nonporous membranes. at the feed interface (phase l/membrane): (V - 131) and at the permeate interface (membrane/phase 2): r11I,'2 m = r111,'2 8 =) aI',2m = aI',2 8exp[-Vi(PRlT - P2)] (V - 132) The activities at the feed interface can be written as (V - 133) while the activities at the permeate interface are Ci,2 m l2 m = Ci,2 8 11,2 8 exp [ -y. (PRIT - P2)] (V - 134) 1 If the solubility constant Kj is defined as the ratio of the activity coefficients, we can write: - - mK _ Yi,1 8 and Ki2 = _Y1_2,-8 (V - 135) 1,1 Yi,1 , Yi,2 m Furthermore, if it is assumed that the diffusion coefficient is concentration-independent, Fick's law (eq. V - 83) can be integrated across the membrane to give = - tJ1 D' (CI',2 m - CI,'l m) (V - 136) ---1.. After substitution of eqs. V - 133, V - 134 and V - 135 into eq. V - 136 one arrives at J = tDj (KI C'lm - K2 c'I2,mexp[_-V-,j,-,(~P...:R.I_-T----,P2=-)]) (V - 137) 1 I, I, I,
190 CHAPTER V and if (Xi = Kj ,2 / Kj,l and Pj = Kj • Dj , then eq. V - 137 converts into = el Pi (c'lS _ (X. c'2 s e x p [ - V d PRlT - P2)]) (V - 138) 1 1, 1 I, Eq. V - 138 is the basic equation used to compare various membrane processes when transport occurs by diffusion [22]. The phases involved in such processes are summarised in table V - 9. TABLE V - 9. Phases involved in diffusion controlled membrane processes Process Phase 1 Phase 2 hyperfiltration LL dialysis gas separation LL pervaporation GG LG V . 6.1 Hyperfiltration Hyperfiltration is nonnally used with aqueous solutions containing a low molecular weight solute, which is often a salt. It can also be used for aqueous solutions containing very small amounts of organic solutes. This process involves the application of pressure to the liquid feed mixture as driving force, the total flux being given by the sum of the water flux Jw and the solute flux Js' With highly selective membranes the solute flux can be neglected (in fact even with less selective membranes the solvent flux is large compared to the solute flux). (V - 139) since ~1t = RT/Vj • (In cw,2s /cwi) and (Xj =1, the water flux Jw may be written as (V - 140) or (V - 141) For small values of x, the tenn (V - 142) 1 - exp(-x) \"\" -x and since
TRANSPORT IN MEMBRANES 191 eq. V - 140 becomes (V - 143) (V - 144) This equation gives the water flux through a membrane as a function of the pressure difference.This equation can also be written in a simple form as: (V - 145) r. t.with A = Dw' Cw Vw/ RT . A is called the water permeability coefficient. Hyperfiltration membranes are generally not completely semipermeable and a simple equation can also be derived for the solute flux. Thus from eq. V - 137, with (Xj =1, the solute flux Js can be written as - t - )Js = Ds Ks (Css.l - Cs.2 s exp [ - Vs (PIR -T P2 - ~1t)J (V - 146) and since the exponential term is approximately unity (see section V - 6.4), eq. V - 146 becomes (V - 147) or (V - 148) where B = Ds.KJt and is called the solute permeability coefficient. Eq. V - 148 expresses in a simple way how the solute flux in hyperfiltration is proportional to the concentration difference, whereas the water (or solvent) flux is proportional to the applied pressure or effective pressure difference (eq. V - 145). V . 6.2 Dialysis In dialysis, liquid phases containing the same solvent are present on both sides of the membrane in the absence of a pressure difference. The pressure terms can therefore be neglected and the following equation may be obtained from eq. V - 138 if (Xi = 1. Ji = ~i ( ci,l s - Ci.2 s ) (V - 149) or
192 CHAPTER V (V - 150) This simple equation describes the solute flux in dialysis indicating that it is proportional to the concentration difference. Separation arises from differences in permeability coefficients: thus macromolecules have much lower diffusion coefficients and distribution coeffients than low molecular weight components. .Y.......Q.3. Gas permeation Gas permeation or vapour permeation, both the upstream and downstream sides of a membrane consist of a gas or a vapour. However, eq. V - 138 cannot be used directly for gases. The concentration of a gas in a membrane can be written as (V - 151) and combination of eq. V-lSI with eq. V - 138 and (Xi =1 gives (V - 152) It can be seen from this equation that the rate of gas permeation is proportional to the partial pressure difference across the membrane. V . 6.4 Pervaporation Pervaporation is a membrane process in which the feed side is a liquid while the permeate side is a vapour as a result of applying a very low pressure downstream. Hence, on the downstream side P2 ~ 0 (or a2s ~ 0) and the exponential term in eq.V - 137 is equal to unity and can be neglected (LW=105 N/m2, Vi = 10-4 m3/mol, RT= 2500 J/mol ~ exp(-Vi.LW/RT) = 1). If the partial pressure is put equal to the activity, then: (V - 153) and eq. V - 138 becomes (V - 154) eJi = p.1 c·1, I S ( 1 _ _p.12,S_) Pi,! S v .7 Transport in ion-exchange membranes Hyperfiltration can be used for the separation of ions from an aqueous solution. Neutral
TRANSPORT IN MEMBRANES 193 membranes are mainly used for such processes and the transport of ions is determined by their solubility and diffusivity in the membrane (as expressed by the solute permeability coefficient, see eq. V - 148). The driving force for ion transport is the concentration difference, but if charged membranes or ion-exchange membranes are used instead of neutral membranes ion transport is also affected by the presence of the fixed charge. Teorell [24] and Meyer and Sievers [25] have used a fixed charge theory to describe ionic transport through these type of systems. This theory is based on two principles: the Nernst-Planck equation and Donnan equilibrium. If an ion-exchange membrane in contact with an ionic solution is considered, then ions with the same charge as the fixed ions in the membrane are excluded and cannot pass through the membrane. This effect is known as Donnan exclusion and can be described by equilibrium thermodynamics which allow the chemical potential of the ionic component in the two phases present to be calculated when an ionic solution is in equilibrium with an ionic membrane. Thus, in the ionic solution itself: (V - 155) where activities are better employed than concentrations because electrolyte solutions generally behave non-ideal (Ideal behaviour may be assumed at very low concentrations.) The activity of a cation or anion is expressed here as the product of the molal concentration m and the activity coefficient y. In the membrane: (V - 156) Quantities with the subscript m refer to the membrane phase. At equilibrium the electrochemical potentials in both phases are equal, thus If the reference states for both phases are also assumed to be equal (/lOj = /lOjm), the following equation may be obtained, with Edon = 'I'm - '1'. Emu = R fTF In (Yi,m mi,m) (V - 157) Zj Yj mj (V - 158) or (V - 159) or , for the case of dilute solutions where aj = Cj
194 CHAPTER V (V - 160) This equation enables some simple calculations to be undertaken. For a given monovalent ionic solute at a concentration difference *o2f 9180),/ the equilibrium potential difference established at the interface is Eden = [(8.314 (96500)] In (1/10) = - 59 mY. In fact an additional term 1t ,Vi' i.e. the swelling pressure originating from the swelling of the crosslinked polymeric network, has to be added to the right- hand side of eq V - 160. This term, however, has little influence on the ionic distribution. The swelling pressure is mainly determined by the concentration of the fixed charge (ion-exchange capacity). The Donnan potential gives the potential build-up at the membrane-solution interface, which is determined by the ionic distribution as shown schematically in figure V - 24 [26]. Indeed, this ionic distribution largely determines the transport of charged molecules. In this example depicted in figure V - 24, the anions are repelled from the interface, since they have the same charge as the fixed charge on the ion-exchange membrane. + + + + + membrane solution phase Figure V - 24. Schematic drawing of the ionic distribution at the membrane-solution interface (membrane contains fixed negatively charged groups). Let us now consider an ion-exchange membrane with a fixed negative charge CR-) with Na+ as the counterion placed in contact with a dilute sodium chloride (NaCl) solution, as shown in figure V - 25. If it is assumed that the solution behaves ideal, the activities can be put equal to the concentrations (ai = ci ). The Na+ and Cl- ions and the water molecules can freely diffuse from the solution to the membrane phase, although the Na+ ions can only diffuse in combination with a Cl- ion. At equilibrium, an equal number of Na+ and Cl- ions will have diffused to either side.
TRANSPORT IN MEMBRANES 195 Na+ ~Na+ cr ~CI Hp ~Hp R- membrane solution phase Figure V - 25. Donnan equilibrium established when an ionic membrane with a fixed negative charge is placed in contact with aqueous NaCI solution. This means that (V - 161) where the superscript m refers to the membrane phase. Because of electrical neutrality (V - 162) Combination of eqs. V - 161 and V - 162 gives (V - 163) and if c [Na+J = [cCi-] , eq. V - 163 becomes or (V - 164) (V - 165) J1[ccd=A-V CK]m + [ca-)ID [ca-)ID For a dilute solution eq. V - 165 reduces to (V - 166) This equation gives the ionic or 'Donnan equilibrium' of charged solutes in the presence of a charged membrane (or charged macromolecules) possessing a fixed charge density R-.
196 CHAPTER V If the concentration in the feed is low and the concentration of the fixed charge (R- in the above example) is high, the Donnan exclusion is very effective. However, with increasing feed concentration, this exclusion becomes less effective. For instance, in the case of brackish water with a concentration of 590 ppm NaCI (= 0.01 eq/l = 10-5 eq/ml) and a membrane with a wet-charge density of = 2 10-3 eq/ml, the co-ion (chloride) concentration in the membrane as estimated from eq. V - 166 will be approximately 5 10-8 eq/ml. This example indicates that the concentration of the co-ion in the membrane is very low and is strongly dependent both on the feed concentration and on the fixed charge density in the membrane. Ionic solutions do not generally behave in an ideal manner and eq. V - 166 must therefore include activity coefficients in order to correct for the non-ideality. Introducing the mean ionic activity coefficients y± (for a univalent cation and anion y± = (y+ . y-)0.5, where y+ and y- are the activity coefficients of the cation and anion, respectively), eq. V - 166 becomes (V - 167) Ion-exchange membranes are frequently used in combination with an electrical potential difference (as in electrodialysis, for example see chapter VI, where electrical driven membrane processes are involved). Two forces now act on the ionic solutes; a concentration difference and an electrical potential difference. Under these circumstances transport of an ion can be described by a combination of these two processes, i.e. a Fickian diffusion and an ionic conductance. The resulting equation is known as the Nemst-Planck equation: J. = - D.Id.lxk - z· F ,.,.,. c·ld.d.xE. (V - 168) 1 1 ••.. Processes driven by electric potentials will be described further in chapters VI and VII. v .8 Literature 1. Prigogine, I., Thermodynamics of irreversible processes, Thomas Springfield, Illinois, 1955 2. Lakshminarayanaiah, N., Transport phenomena in membranes, Academic Press, Orlando, USA,. 1969 3. Katchalsky, A., and Curran P.F., Non-equilibrium processes in biophysics, Harvard University Press, 1965 4. Mason, E.A., and Malinauskas, A.P., Gas transport in porous media: The dusty gas model, Elsevier, Amsterdam, 1983 5. Kedem,O., and Katchalsky, A.,Biochim. Biophys. Acta., 27 (1958) 229 6. Spiegler, K.S., and Kedem, 0., Desalination, 1 (1966) 311 7. Nakao, S.L, and Kimura, S.1., J. Chem. Eng. Japan, 14 (1981) 32 8. Jonsson, G., and Boessen, C.E., Proc. 6th. Symp. Fresh Water from the Sea,
TRANSPORT IN MEMBRANES 197 1978, Vol. 3, p. 157 9. Vieth, W.R., Howell, J.M., and Hsieh, R.J.,J. Membr., Sci, 1 (1976) 177 10. Fujita, H., Fortschr. Hochpolym. Forsch, 3 (1961) 1 11. Flory, P.J., Principles ofPolymer Chemistry, Cornell Univ. Press, Ithaca, 1953 12. Park, G.S.,: 'Transport in polymers', in: Bungay, P.M., Lonsdale, H.K, and de Pinho (eds.), Synthetic Membranes: Science, Engineering, and Applications, Reidel Publishing Company, Dordrecht, 1986, p. 57 13. Brown, W.R, and Park, G.S., 1. Paint. Technol, 42 (1970) 16 14. Baker, RW., and Blume, I., Chemtech., 16 (1986) 232 15. Me Call, D.W.,J. Polym. Sci., 16 (1975) 151 16. Daynes, H.A., Proc. Roy. Soc., A., 97 (1920) 286 17. Blume, I., Mulder, M.H.V., and Smolders, c.A., to be published 18. Chandler, H.W., and Henley, E.J., AIChEJ, 7 (1961) 295 19. Vrentas, J.S., and Duda, J.L., J. Polym. Sci., Polym. Phys., 15 (1977) 403 20. Vrentas, J.S., and Duda, J.L., J. Polym. Sci., Polym. Phys., 15 (1977) 417 21. Mulder, M.H.V., Franken, A.C.M., and Smolders, C.A.,J. Membr. Sci., 22, (1985), 155 22. Lee, C.H., J. Appl. Polym. Sci., 19 (1975) 83 23. Krishna, R., and Wesselingh, H.J., to be published 24. Teorell, T., Proc. Soc. Exp. Bioi. Med., 33 (1935) 282 25. Meyer, KH., and Sievers, J-F., Helv. Chim. Acta, 19 (1936) 649 26. Helfferieh, F., lon-Exchange, McGraw-Hill, New York, 1962 27. Mulder, M.H.V.; 'Thermodynamics principles of Pervaporation' in R.Y.M. Huang (ed.), Pervaporation Membrane Separation Processes, Elsevier, Amsterdam, 1991, Chapter 4. 28. Frisch, H.L., J. Phys. Chem., 62 (1957) 93 29. Kokes, R.J., and Long, F.A., J. Am. Chem. Soc., 75 (1953) 6142 30. Cohen, M.H., and Turnbull, D., J. Chem. Phys., 31 (1959) 1164 31. Simha, R., and Boyer, R.F., J. Chem., Phys., 37 (1962) 1003 32. Bondi, A., J. Phys. Chem., 68 (1964) 411 33. Sugden, S., J. Chem. Soc., (1927), 1786 34. Muruganandam, N., Koros, W.J., Paul, D.R.,J. Polym. Sci. Polym. Phys., 25 (1987) 1999 35. Barbari, T., Koros, W.J., and Paul, D.R., J. Polym. Sci. Polym. Phys., 26 (1988) 709 36. Min, K.E., and Paul, D.R, J. Polym. Sci. Polym. Phys., 26 (1988) 1021 37. Tanaka, K, Kita, H., Okamoto, K, Nakamura, A., Kusuki, Y., J. Membr. Sci, 47 (1989) 203 38. Smit, E., Mulder, M.H.V., and Smolders, C.A., to be published 39. Zimm, B.H., and Lundberg, J.L., J. Phys. Chem., 60 (1956) 425 40. Crank, J., The mathematics of diffusion, Clarendon Press, Oxford, 1975 41. Bitter, J.G.A., Desalination, 51 (1984) 19 42. Paul, D.R., and Koros, J.W., J. Polym. Sci. Polym. Phys, 14 (1976) 675 43. Petropolous, I.H., J. Polym. Sci. A-2, 8 (1970) 1797 44. Staverman, A.J., Rec. Trav. Chim, 70 (1951) 344 45. Blume, I., to be published
VI MEMBRANE PROCESSES VI . 1 Introduction All membrane processes have the common feature that separation is achieved via a membrane. The membrane can be considered to be a permselective barrier existing between two homogeneous phases. Transport through the membrane takes place when a driving force is applied to the components in phase 1. In most the membrane processes the driving force is a pressure difference or a concentration (or activity) difference across the membrane. Parameters such as pressure, concentration (or activity) and even temperature may be included in one parameter, the chemical potential 11. 11 = f (T, P, a or c) (VI - 1) At constant temperature, the chemical potential of component i in a mixture is given by (VI - 2) where Iloi is the chemical potential of 1 mol of pure substance at a pressure P and temperature T. For pure components the activity is unity, i.e. a = 1, but for liquid mixtures the activity is given by (VI - 3) where xi is the mole fraction and 'Y i is the activity coefficient. For ideal mixtures the activity coefficient is unity, i.e. 'Y i = 1, so that the activity is equal to the mole fraction, i.e. aj = xi. However, since many non-aqueous mixtures are non-ideal, activities rather than concentrations should be used. For gases, the chemical potential is given by (VI - 4) where Pi is the partial pressure. Because (VI - 5) eq. VI - 4 becomes (VI - 6) Ili = Iloi + RT In Xj + RT lnP Gases deviate from ideality with increasing pressure and in fact fugacities should be used instead of partial pressures. Eq. VI - 4 then becomes 198
MEMBRANE PROCESSES 199 (VI -7) where fi is the fugacity. Table VI - 1 gives an example of the difference between pressure and fugacity for carbon dioxide at various pressures at 300 K. It can be seen that for high pressures the discrepancy between pressure and fugacity can be quite large. TABLE VI - 1. The fugacity of carbon dioxide at 300 K [1] pressure (P) fugacity (t) fIP (bar) (bar) 1 0.995 0.995 5 4.9 0.976 25 22.0 0.880 50 38.1 0.761 60 42.8 0.713 Another type of driving force is the electrical potential difference across the membrane. However, this driving force only influences the transport of charged particles or molecules. The membrane processes discussed in this chapter may be classified according to their driving forces. Such a classification is given in table VI - 2. TABLE VI - 2. Classification of membrane processes according to their driving forces pressure concentration temperature electrical difference (activity) difference difference potential difference microfiltration pcrvaporation thermo-osmosis electrodialysis ultrafiltration gas separation membrane distillation electro-osmosis hyperfiltration dialysis piezodialysis liquid membranes Before describing these various processes an introduction is given on osmotic phenomena, because such phenomena are very important in membrane processes especially in pressure-driven processes. VI . 2 Osmosis An osmotic pressure arises when two solutions of different concentration (or a pure
200 CHAPTER VI permeable to the solvent but impermeable to the solute. This situation is illustrated schematically in figure VI - la. Here the membrane separates two liquid phases: a concentrated phase 1 and a dilute phase 2. rnernbmne solvent Figure VI - 1. Schematic illustration of osmotic processes. Under isothermal conditions the chemical potential of the solvent in the concentrated phase (phase 1) is given by (VI - 8) while the chemical potential of the solvent in the dilute phase (phase 2) is given by (VI - 9) The solvent molecules in the dilute phase have a higher chemical potential than those in the concentrated phase (because ai.2 or In ai.2 is greater than au or In aU)' This chemical potential difference causes a flow of solvent molecules from the dilute phase (high chemical potential) to the concentrated phase (low chemical potential). This is shown in figure VI - 1b. This process continues until osmotic equilibrium has been reached, i.e. when the chemical potentials of the solvent molecules in both phases are equal (see figure VI - lc): Ili.l = 1li,2 (VI - 10) Combination of eqs. VI - 8, VI - 9 and VI - 10 gives, (VI - 11) RT (In !!j.1 - In ai.2) = (PI - P2) Vi =~1t . Vi This hydrodynamic pressure difference (PI - P2) is called the osmotic pressure difference ~1t (~1t = 1t1 - 1t2)' When only pure solvent is situated on one side of the membrane, i.e.ai = 1, then eq. VI - 11 becomes
MEMBRANE PROCESSES 201 1t = - RVT I n a · 2 (VI - 12) 1 I, For very low solute concentrations (y i ~ 1) eq. VI - 12 can be simplified further by applying Raoult's law: In a·1 = In y.1 x·1 \"\" In x·1 \"\" In (1 - xJ·) =- x·J (VI - 13) 1t (VI -14) For a dilute solution, Xj \"\" n/ni and (VI-IS) or 1t V = n·J RT (VI - 16) 1 (VI - 17) Because Vi'\" V (for dilute solutions) and nj N = Cj / M, then This simple relationship between the osmotic pressure 1t and the solute concentration Cj' is called the van 't Hoff equation, It can be seen that the osmotic pressure is proportional to the concentration and inversely proportional to the molecular weight. If the solute dissociates (as for instance in salts) or associates, eq. VI - 17 must be modified. When dissociation occurs the number of moles increases and hence the osmotic pressure increases, whereas in the case of association the number of moles decreases as does the osmotic pressure. Some examples calculated on the basis of eq. VI - 17 are given in table 3. Table VI - 3. Calculation of osmotic pressures of some aqueous solutions The osmotic pressures of aqueous solutions containing 3% NaCl (M = 58.45) by weight, 3% albumin (M = 65.000) by weight and a suspension containing 30 gil of a solid (where the particle weight is 1 ng = 10-9 g) can be calculated at a temperature 25\"C using eq. VI - 10, viz. NaCl: 7t = 2 * (30/58.45) * 8.31 * 298.2 = 2.54106 Pascal = 25.4 Bar Albumin: 7t = (30/65000) * 8.31 * 298.2 = 1.14103 Pascal=O.OI Bar Suspension 7t = (30109/6.23 1023 ) * 8.31 * 298.2 = 1.19 10-7 Pascal = 10-12 Bar
202 CHAPTER VI These examples show that the osmotic pressure can be neglected in microfiltration and ultrafiltration applications, whereas it has to be taken into account in hyperfiltration. Substantial deviations from van 't Hoffs law occur with macromolecular solutions as will be described in chapter VII. VI . 3 Pressure driven membrane processes VI . 3.1 Introduction Various pressure-driven membrane processes can be used to concentrate or purify a dilute (aqueous) solution. The particle or molecular size and chemical properties of the solute determine the structure, i.e. pore size and pore size distribution, necessary for the membrane employed. Various processes can be distinguished related to the particle size of the solute and consequently to membrane structure. These processes are microfiltration, ultrafiltration or hyperfiltration (reverse osmosis). The principle of the three processes is illustrated in figure VI - 2. microfiltration solvent o solute (low mol. weight) • solute (high mol. weight) ® panicle ultrafiltration feed permeate hyperfiltration ~p Figure VI - 2. Schematic representation of microfiltration, ultrafiltration and hyperfiltration. Because of a driving force, i.e. the applied pressure, the solvent and various solute molecules permeate through the membrane, whereas other molecules or particles are rejected to various extents. As we go from microfiltration through ultrafiltration to hyperfiltration, the size (molecular weight) of the particles or molecules separated diminishes and consequently the pore sizes in the membrane must become smaller. This
MEMBRANE PROCESSES 203 implies that the resistance of the membranes to mass transfer increases and hence the applied pressure (driving force) has to be increased. However, no sharp distinction can be drawn between the various processes. A schematic drawing of the separation range involved in these various processes is given in figure VI - 3. .... ultrafiltration .... ...microfiltration .......hyperfiltration 0.1 1.0 10 100 1000 molecular size (nm) Figure VI - 3. Application range of microfiltration, ultrafiltration and hyperfiltration. It is possible to distinguish between the various processes in terms of membrane structure. In the case of microfiltration, the complete membrane thickness contributes towards transport resistance, especially when a symmetrical porous structure is involved. The membrane thickness can extend from lO!-lm to more than 150 !-lm. With ultrafiltration and TABLE VI - 4. Comparison of various pressure driven membrane processes microfiltration ultrafiltration hyperfiltration separation of particles separation of macromolecules separation of low MW (bacteria, yeasts) (proteins) solutes (salt, glucose, lactose) osmotic pressure osmotic pressure osmotic pressure high negligible negligible ('\" 5 - 25 bar) applied pressure low applied pressure low applied pressure high ('\" 1- lObar) ('\" 10 - 60 bar) « 2 bar) symmetric structure asymmetric structure asymmetric structure (not always) thickness of separating thickness of actual separating thickness of actual layer'\" 10 - 150 11m layer '\" 0.1 - 1.0 11m separating layer\", 0.1-1.0 11m separation based on particle separation based on particle size separation based on size differences in solubility and diffusivity
204 CHAPTER VI hyperfiltration, on the other hand, asymmetric membranes are used in which a thin, relatively dense top layer (thickness 0.1-1.0 J.lm) is supported by a porous substructure (thickness\"\" 50-150 J.lm). The hydraulic resistance is almost completely located in the top layer, the sublayer having only a supporting function. The flux through these (and other) membranes is inversely proportional to the (effective) thickness, and because they possess an asymmetrical structure with top layer thicknesses less than 1 J.lm membranes of this type became of commercial interest. A comparison of the various processes is given in table VI - 3. VI . 3.2 Microfiltration Microfiltration is the membrane process which most closely resembles conventional coarse filtration. The pore sizes of microfiltration membranes range from 10 to 0.05 J.lm, making the process suitable for retaining suspensions and emulsions. The volume flow through these microfiltration membranes can be described by Darcy's law, the flux J through the membrane being directly proportional to the applied pressure: J=K L\\P (VI - 18) where the permeability constant K contains structural factors such as the porosity and pore size (pore size distribution). Furthermore, the viscosity of the permeating liquid is also included in this constant. If the membrane consists of straight capillaries, the Hagen- Poiseuille relationship can be used with K \"\" E r2: J (VI-19) When a nodular structure exists, i.e. where these structure is akin to an assembly of spherical particles, the Kozeny-Carman equation can be employed: J E3 L1P (VI - 20) K T\\ S2 (1 - E)2 L1x In both eqs. VI - 19 and VI - 20, the viscosity appears as an inversely proportional parameter. Also both equations relate the volume flow to simple structural parameters such as porosity E and pore radius r. The flux is proportional to the porosity in capillary membranes whereas with nodular types of structure the dependence on porosity is more complex. In order to optimise microfiltration membranes, it is essential to ensure that the structural parameters are such that the (surface) porosity is as high as possible with the pore size distribution as narrow as possible. It should be realised that the convective flow as described by these equations only involves membrane-related parameters and none which apply to the solutes. VI . 3.2.1 Membranes for microfiltration Microfiltration membranes may be prepared from a large number of different materials based on either organic materials (polymers) or inorganic materials (ceramics, metals, glasses). Various techniques can be employed to prepare microfiltration membranes from
MEMBRANE PROCESSES 205 polymeric materials: (a) (b) (c) Figure VI - 4. Polymeric microfiltration membranes: (a) phase inversion; (b) stretching; and (c) track etching - sintering - stretching - track-etching - phase inversion. Such preparation techniques have already been discussed in detail in chapter III. Figure VI - 4 shows SEM micrographs of some characteristic polymeric microfiltration membranes obtained by phase inversion (IV - 4a), stretching (IV - 4b) or track-etching (IV - 4c). Frequently, inorganic membranes are used instead of polymeric membranes because of their outstanding chemical and thermal resistances. In addition, the pore size in these membranes can be better controlled and as a consequence the pore size distribution is generally very narrow (see also chapter IV). Various techniques can be used to prepare ceramic membranes with some important ones being: - sintering - sol/gel process - anodic oxidation
206 CHAPTER VI Two typical structures are depicted in figure VI - 5. (a) (b) Figure VI - 5. Ceramic microfiltration membranes: (a) Anotec®, anodic oxidation (surface); and (b) Ceraver®, sintering (cross section, upper part). These SEM photographs clearly show that the porosity and pore size distribution differ substantially for the various membranes depicted. Table VI - 5 summarises the effect of the preparation method on the porosity and the pore size distribution. Table VI - 5. Porosities and pore size distributions achieved by various preparation methods process porosity pore size distribution sintering low/medium narrow/wide stretching medium/high narrow/wide track-etching low narrow phase inversion high narrow/wide
MEMBRANE PROCESSES 207 These various techniques microfiltration membranes to be made from virtually all kinds of materials of which polymers and ceramics are the most important. Synthetic polymeric membranes can be divided in two classes, i.e. hydrophobic and hydrophilic. Various polymers which yield hydrophobic and hydrophilic membranes are listed below. Ceramic membranes are based mainly on two materials, alumina (AI20 3) and zirconia (Zr02)' However, other materials such as titania (Ti02) can also be used in principle. A number of organic and inorganic materials are listed below: - hydrophobic polymeric membranes polytetrafluoroethylene (PTFE, teflon) poly(vinylidene fluoride) (PVDF) polypropylene (PP) - hydrophilic polymeric membranes cellulose esters polycarbonate (PC) polysulfone/poly(ether sulfone) (PSf/PES) polyimide/poly(ether imide) (PI/PEI) (aliphatic) polyamide (PA) - ceramic membranes alumina (AI20 3) zirconia (Zr02) Other materials such as glass (Si02) and various metals (palladium, tungsten, silver) have also been used for preparing microfiltration membranes. Microfiltration membranes, possessing pores in the range 0.1 - 2 1lID, are relatively easy to characterise (see chapter IV). The main techniques employed are Scanning Electron Microscopy (SEM), bubble-point measurements, mercury porometry and permeation measurements. The main problem encountered when rnicrofiltration is applied (in the laboratory or on an industrial scale) is flux decline. This is caused by concentration polarisation and fouling (the latter being the deposition of solutes inside the pores of the membrane or at the membrane surface). Quite often considerable flux declines can be observed with values for process fluxes approximately 1% of the pure water flux being not unrealistic. These phenomena will be discussed in detail in chapter VII. To reduce fouling as much as possible it is important that careful control is exercised over the mode of process operation. Basically, two process modes exist, i.e. dead-end and cross-flow filtration (see also chapter VIII). In dead-end filtration the feed flow is perpendicular to the membrane surface, so that the retained particles accumulate and form a type of a cake layer at the membrane surface. The thickness of the cake increases with filtration time and consequently the permeation rate decreases with increasing cake layer thickness. In cross- flow filtration the feed flow is along the membrane surface, so that part of the retained solutes accumulate. A schematic drawing of these processes is shown in figure VI - 6. Adsorption phenomena may also play an important role in fouling and hence it is important to the select an appropriate membrane material. Hydrophobic materials of the type mentioned above have a larger tendency to foul in general, especially in the case of proteins. Furthermore, such hydrophobic materials (e.g. polytetrafluoroethylene) are not wetted by water and no water will flow through the membrane at normal applied pressures. This non-wettability is another disadvantage and such membranes have to be pretreat, for example with alcohol, prior to use with aqueous solutions. Flux decline still occurs despite a proper choice of the process mode since it is an implicit part of the process and the membranes must be cleaned periodically. This implies that the choice of the membrane material is also important with respect to its stability relative to the
208 CHAPTER VI feed feed retentate ~-------~ permeate +permeate Idead-end I Icross-flow Figure VI - 6. Schematic representation of dead-end filtration and cross-flow filtration. cleaning procedure. An example of a chemical which is frequently used as a cleaning agent is active chlorine, towards which many polymers are not stable. Furthermore, resistance over a wide pH range is another requirement in chemical stability. Other applications, especially in biotechnology, require stability against steam sterilisation and this must extend over the whole module including the housing and potting materials. All these examples clearly indicate that not only is the membrane performance important in microfiltration but particularly the chemical and thermal resistance of the materials used. VI . 3.2.2 Industrial applications Microfiltration is used in a wide variety of industrial applications where particles of a size > 0.1 Ilm, have to be retained from a liquid. The main applications are in small-scale (analytical) processes in all kinds of laboratories. One of the main industrial applications is the sterilisation and clarification of all kinds of beverages and pharmaceuticals in the food and pharmaceutical industries. This can be done at any temperature, even at low temperatures. This process is also used to remove particles during the processing of ultrapure water in the semiconductor industry. New fields of application are biotechnology and biomedical technology. In biotechnology, microfiltration is especially suitable in cell harvesting and as a part of a membrane bioreactor (involving a combination of biological conversion and separation). In the biomedical field, plasmapheresis which involves the separation of plasma with its value products from blood cells appears to have an enormous potential. A number of applications are summarised below [2]: - cold sterilisation of beverages and pharmaceuticals - cell harvesting - clarification offruitjuice, wine and beer - ultrapure water in the semiconductor industry - metal recovery as colloidal oxides or hydroxides - waste-water treatment - continuous fermentation - separation of oil-water emulsions - dehydration of latices
MEMBRANE PROCESSES 209 VI . 3.2.3 Summary of microfiltration membranes: (a)symmetric porous thickness: ,., 10 -150 11m pore sizes: ,., 0.05 - 10 11m driving force: separation principle: pressure « 2 bar) membrane material: main applications: sieving mechanism polymeric, ceramic - analytical applications - sterilisation (food, pharmaceuticals) - Ultrapure water (semiconductors) - clarification (beverages) - cell harvesting and membrane bioreactor (biotechnology) - plasmapheresis (medical) VI . 3.3 Ultrafiltration Ultrafiltration is a membrane process whose nature lies between hyperfiltration (or reverse osmosis) and microfiltration. The pore sizes of the membranes used range from 0.05 11m (on the rnicrofiltration side) to 1 nm (on the hyperfiltration side). Ultrafiltration is typically used to retain macromolecules from a solution, the lower limit being solutes with molecular weights of a few thousand Daltons. A range of solutes with molecular weights extending from a few hundred up to few thousand Daltons may be treated by a process in between ultrafiltration and hyperfiltration which is often referred to as nanofiltration. Ultrafiltration and rnicrofiltration membranes can both be considered as porous membranes where rejection is determined mainly by the size and shape of the solutes relative to the pore size in the membrane and where the transport of solvent is directly proportional to the applied pressure. Such convective solvent flow through a porous membrane can be described by the Kozeny-Carman equation (see eq. VI - 16) for example. In fact both microfiltration and ultrafiltration involve similar membrane processes based on the same separation principle. However, an important difference is that ultrafiltration membranes have an asymmetric structure where the hydrodynamic resistance is mainly determined within a small part of the total membrane thickness, whereas in rnicrofiltration virtually the whole of the membrane thickness contributes towards the hydrodynamic resistance. The top layer thickness in an ultrafiltration membrane is generally less than 1 11m with figure VI - 7 illustrating an example of an asymmetric polysulfone membrane. The flux through an ultrafiltration membrane can be described in the same way as for microfiltration membranes, being directly proportional to the applied pressure: J=K ~p (V - 18) The permeability constant K includes all kinds of structural factors. The value of this constant K for ultrafiltration membranes is much smaller than for microfiltration membranes, being of the order of 0. 1 m3/m2.day.bar for dense membranes up to about 10 m3/m2.day.bar for the more open membranes.
210 CHAPTER VI Figure VI - 7. A scanning electron micrograph showing the cross-section of an ultrafiltration polysulfone membrane (magnification: 10,000 x). VI . 3.3.1 Membranes for ultrafiltration Most of ultrafiltration membranes used commercially these days are prepared from polymeric materials by a phase inversion process. Some of these materials are listed below: - polysulfone/poly(ether sulfone)/sulfonated polysulfone - poly(vinylidene fluoride) - polyacrylonitrile (and related block-copolymers) - cellulosics (e.g. cellulose acetate) - polyimide/poly(ether imide) - aliphatic polyamides In addition to such polymeric materials, inorganic (ceramic) materials have also been used for ultrafiltration membranes. Figure VI - 8 shows an Al20 3 membrane prepared via a sol- gel technique. Since the lower limit for preparing porous membranes by sintering is about 0.1 11m pore diameter, this technique cannot be used to prepare ultrafiltration membranes. Such sintered porous structures can be used as the sublayer for composite ultrafiltration membranes, a technique frequently used in the preparation of the ceramic ultrafiltration membranes. On the other hand, ultrafiltration membranes themselves are often used as sublayers in composite membranes for reverse osmosis, gas separation and pervaporation. Ultrafiltration is often used for the fractionation of macromolecules where large molecules have to be retained by the membrane while small molecules (and the solvent) should permeate through freely . In order to choose a suitable membrane, manufacturers often used the concept of 'cut-off (see chapter IV). However, it should be realised that molecular weight is not the only criterion that determines selectivity. An important
MEMBRANE PROCESSES 211 Figure VI - 8. SEM photograph of an inorganic Al20 3 membrane. factor with a large influence on cut-off measurements (or on separation in general) is the occurrence of concentration polarisation. For instance a membrane with a cut-off of 40,000 is almost completely permeable towards cytochrome-c (Mw: 14,400). However, in a mixture of cytochrome-c with bovine serum albumin (Mw: 67,000), both albumin and a large portion of the cytochrome-c are rejected. This is because of concentration polarisation. The rejected albumin molecules build up a secondary layer at the membrane surface which acts as a dynamic membrane capable of retaining cytochrome-c molecules. Furthermore, different types of solute, e.g. globular proteins or linear macromolecules such as polyethylene glycol and dextran, possess different rejection characteristics. Thus, a cut-off value is not an intrinsic membrane parameter and corrections must be made for the occurrence of concentration polarisation and the fact that most polymers possess a molecular weight distribution. This has been described in more detail in chapter IV. There a number of other techniques besides cut-off measurements for characterising ultrafiltration membranes. However, typical methods for microfiltration membranes, such as mercury intrusion or scanning electron microscopy cannot be used for the characterisation of ultrafiltration membranes. For this reason, other techniques have been developed such as thermoporometry and permporometry as discussed in chapter IV. Other more general techniques which are applicable are gas adsorption-desorption, permeability measurements and 'modified cut-off measurements. An important point which must be considered is that as in microfiltration the process performance is not equal to the intrinsic membrane properties in actual separations. The reason for this is again the occurrence of concentration polarisation and fouling. The macromolecular solute retained by the membrane accumulates at the surface of the membrane resulting in a concentration build-up. At steady state, the convective flow of the solute to the membrane is equal to the diffusional back-flow from the membrane to the bulk. Further pressure increase will not result in an increase in flux because the resistance of the boundary layer has increased (see chapter VII) so that a limiting flux value (Joo) is attained (see figure VI - 9). As in microfiltration, these boundary layer phenomena mainly determine the process pelformance. Thus, intrinsic properties are not all that important in membrane development, but rather its resistance to the various kinds of chemicals necessary for cleaning. The number of membrane applications increase as they become
212 CHAPTER VI more resistant to higher temperatures (> 100ce), to a wide range of pH (1 to 14) and to organic solvents. Furthermore, as in microfiltration , module and system design are very important for reducing fouling as much as possible at a minimal cost. pure water J --------s-o-lu-t-io-n-------J= Figure VI - 9. Schematic drawing of the relationship between flux and applied pressure in ultrafiltration. VI . 3.3.2 Applications Ultrafiltration is used over a wide field of applications involving situations where high molecular components have to be separated from low molecular components. Applications can be found in fields such as the food and dairy industry, pharmaceutical industry, textile industry, chemical industry, metallurgy, paper industry, and leather industry [1,3]. Various applications in the food and dairy industry are the concentration of milk and cheese making, the recovery of whey proteins, the recovery of potato starch and proteins, the concentration of egg products, and the clarification of fruit juices and alcoholic beverages. Ultrafiltration membranes have been used up until now for aqueous solutions, but a new and developing field is in non-aqueous applications. For these latter applications, (new) chemical resistant membranes must be developed from more resistant polymers. Inorganic membranes can also be used in this field. VI. 3.3.3 Summary of ultrafiltration membranes: asymmetric porous thickness: '\" 150!lm pore sizes: '\" 1- 100 nm driving force: pressure (1 - 10 bar) separation principle: sieving mechanism membrane material: polymer (e.g. polysulfone, polyacrylonitrile) main applications: ceramic (e.g. zirconium oxide, aluminium oxide) - dairy (milk, whey, cheese making) - food (potato starch and proteins) - metallurgy (oil-water emulsions, electropaint recovery) - textile (indigo) - pharmaceutical (enzymes, antibiotics, pyrogens)
MEMBRANE PROCESSES 213 VI . 3.4 Hyperfiltration Hyperfiltration is used when low molecular weight solutes such as inorganic salts or small organic molecules such as glucose have to be separated from a solvent. The difference between ultrafiltration and hyperfiltration lies in the size of the solute. Consequently, denser membranes are required with a much higher hydrodynamic resistance. Such low molecular solutes would pass freely through ultrafiltration membranes. In fact, hyperfiltration membranes can be considered as being intermediate between open porous types of membrane (microfiltration/ultrafiltration) and dense nonporous membranes (pervaporation/gas separation). Because of their higher membrane resistance, a much higher pressure must be applied to force the same amount of solvent through the membrane. The solutions containing the low molecular weight solutes have a much higher osmotic pressure than the macromolecular solutions used in ultrafiltration. The osmotic pressure of seawater, for example, is about 25 bar. ,1p o ~----~~------------ Iif ,1P < tm Ic:::::::> Jw Jw ~ lif,1P >,11t I Figure VI - 10. Schematic drawing of water flow (Jw) as a function of applied pressure (,1P). Figure VI - 10 presents a schematic drawing of a membrane separating pure water from a salt solution. The membrane is permeable to the solvent (water) but not to the solute (salt). In order to allow water to pass through the membrane, the applied pressure must be higher than the osmotic pressure. As can be seen from figure VI - 10, water flows from the dilute solution (pure water) to the concentrated solution if the applied pressure is smaller than the osmotic pressure. When the applied pressure is higher than the osmotic pressure water flows from the concentrated solution to the dilute solution (see also figure VI - I). The effective water flow can be represented by eq. VI - 17 if it is assumed that no solute permeates through the membrane: (VI - 21) In practice, the membrane may be a little permeable to low molecular solutes and hence the real osmotic pressure difference across the membrane is not ,11t but cr ,11t, where cr is the reflection coefficient of the membrane towards that particular solute (see also chapter V).
214 CHAPTER VI When R < 100%, then cr < 1 and eq. VI - 21 now becomes (VI - 22) In the description considered here, we assume that the solute is completely retained by the membrane. The water permeability coefficient A (also defined as the hydrodynamic permeability coefficient) is a constant for a given membrane and contains the following parameters (see also chapter V - 6.1). A = Dw Cw Vw (VI - 23) R T i1x The value of A, which is a function of the distribution coefficient (solubility) and the diffusivity, lies roughly in the range 5 .10-3 - 5.10-5 m3. m-2.h-l .barl , with the lowest value being observed for denser membranes. The solute flux can be described by (VI - 24) where B is the solute permeability coefficient and i1cs the solute concentration difference across the membrane (i1cs = cf - cp). The value of B lies in the range 5 .10-3 - 104 m .h-l , with the lowest value for high rejection membranes. B = Ds Ks (VI - 25) ;1x Eq. VI - 25 also indicates that the solute permeability coefficient B is a function of the diffusivity and the distribution coefficient (solubility). From eq. VI - 21 it can be seen that when the applied pressure is increased the water flux increases linearly. The solute flux (eq. VI - 24) is hardly affected by the pressure difference and is only determined by the concentration difference across the membrane. The selectivity of a membrane for a given solute is expressed by the retention coefficient or rejection coefficient R: R = Cf - cp = 1 _ cp (VI - 26) Cf Cf Consequently, as the pressure increases the selectivity also increases because the solute concentration in the permeate decreases. The limiting case ~ax is reached as i1p => 00. With cp = J/Jw and combining eqs. VI - 21, VI - 24 and VI - 26, the rejection coefficient can be written as: R A (i1P - i11t) (VI - 27) A (i1P - i11t) + B Eq. VI - 27 is very illustrative since the only variable which appears in this equation is &, assuming that the constants A and B are independent of the pressure (see also chapter
MEMBRANE PROCESSES 215 VIII). The pressures used in hyperfiltration range from 20 to 100 bar, which are much higher than those used in ultrafiltration. In contrast to ultrafiltration and microfiltration, the choice of material directly influences the separation efficiency through the constants A and B (see eq. VI - 27). In simple terms, this means that the constant A must be as high as possible whereas the constant B must be as low as possible to obtain an efficient separation. In other words, the membrane (material) must have a high affinity for the solvent (mostly water) and a low affinity for the solute. This implies that the choice of material is very important because it determines the intrinsic membrane properties. The difference to ultrafiltration/microfiltration, where the dimensions of the pores in the material determine the separation properties and the choice is mainly based upon chemical resistance, is obvious. VI . 3.4.1 Membranes for hyperfiItration The flux through the membrane is as important as its selectivity towards various kinds of solute. When a given material has been selected on the basis of its intrinsic separation properties, the flux through the membrane prepared from this material can be improved by reducing its thickness. The flux is approximately inversely proportional to the membrane thickness and for this reason most hyperfiltration membranes have an asymmetric structure with a thin dense top layer (thickness :::; 1 Ilffi) supported by a porous sublayer (thickness'\" 50 - 150 ).1m), the resistance towards transport being determined mainly by the dense top layer. Two different types of membrane with an asymmetric structure can be distinguished: i) (integral) asymmetric membranes; and ii) composite membranes. In integral asymmetric membranes, both top layer and the sublayer consist of the same material. These membranes are prepared by phase inversion techniques. For this reason it is essential that the polymeric material from which the membrane it to be prepared is soluble in a solvent or a solvent mixture. Because most polymers are soluble in one or more solvents, asymmetric membranes can be prepared from almost any material. However, this certainly does not imply that all such membranes are suitable for every hyperfiltration application because the material constants A and B must have optimal values for a given application. Thus for aqueous applications, e.g. the desalination of seawater and brackish water, hydrophilic materials should be used (high A value) with a low solute permeability. An important class of asymmetric hyperfiltration membrane prepared by phase inversion are the cellulose esters, especially cellulose diacetate and cellulose triacetate. These materials are very suitable for desalination because of their high permeability towards water in combination with a (very) low solubility towards the salt. However, although the properties of membranes prepared from these materials are very good, their stability against chemicals, temperature and bacteria is very poor. Typical operation conditions of such membranes are over the pH range 5 to 7 and at a temperature below 30°C, thus avoiding hydrolysis of the polymer. The extent of this hydrolysis decreases as the degree of acetylation increases, and for this reason cellulose diacetate is less resistant than cellulose triacetate. Biological degradation is also a severe problem whilst another limitation of cellulose acetate membranes is their rather poor selectivity towards small organic molecules other than carbohydrates such as glucose or sucrose. Other materials that have been used frequently for hyperfiitration membranes are aromatic polyamides. These material also show high selectivities towards salts but their water flux is somewhat lower. Polyamides can be used over a wider pH range, approximately from 5 - 9. The main drawback of polyamides (or of polymers with an amide group -NH-CO in general) is their susceptibility against free chlorine Cl2 which causes degradation of the amide group. Asymmetric membrane as well as symmetric
216 CHAPTER VI membranes have been prepared from these polymers by melt or dry spinning to obtain hollow fibers with very small dimensions (outside diameters of such hollow fibers < 100 /lm). The membrane thickness of these fibers is about \"\" 20 /lm with the result that the permeation rate has decreased dramatically. However, this effect is counteracted by the extremely high membrane surface area in a given volume element, with values up to 30,000 m2/m3 (see also chapter VIII). A third class of material that have been used are the polybenzimidazoles, polybenzimidazolones, polyamidehydrazide and polyimides. The chemical structures of these materials have been given in chapter II. Composite membranes constitute the second type of structure frequently used in hyperfiltration. In such membranes the top layer and sublayer are composed of different polymeric materials so that each layer can be optimised separately. The first stage in manufacturing a composite membrane is the preparation of the porous sublayer. Important criteria for this sublayer are surface porosity and pore size distribution and asymmetric ultrafiltration membranes are often used. Different methods have b.~en employed for placing a thin dense layer on top of this sublayer; - dip coating - in-situ polymerisation - interfacial polymerisation - plasma polymerisation These various methods have been discussed in chapter III. Since hyperfiltration membranes may be considered as intermediate between porous ultrafiltration membranes and very dense nonporous pervaporation/gas separation membranes, it is not necessary that their structure to be as dense as for pervaporation/gas separation. Most composite hyperfiltration membranes are prepared by interfacial polymerisation (see chapter III - 6) in which two very reactive bifunctional monomers (e.g. a di-acid chloride and a di-amine) are allowed to react with each other at a water/organic solvent interface. Examples of monomers used for interfacial polymerisation are given in table VI - 6 (see also table III - 1). TABLE VI - 6. Example of monomers used for interfacial polymerisation piperazine m-terephthaloyl acid chloride polyamide product water-soluble organic solvent- monomer soluble monomer VI . 3.4.2 Auulications Hyperfiltration can be used in principle for a wide range of applications, which may be roughly classified as solvent purification (where the permeate is the product) and solute concentration (where the feed is the product). Most of applications are in the purification of water, mainly the desalination of
MEMBRANE PROCESSES 217 brackish and especially seawater to produce potable water [1,4]. The amount of salt present in brackish water is between 1000-5000 ppm, whereas in seawater the salt concentration is about 35,000 ppm. Another important application is in the production of ultrapure water for the semiconductor industry. Hyperfiltration is used as a concentration step particularly in the food industry (concentration of fruit juice, sugar, coffee), the galvanic industry (concentration of waste streams) and the dairy industry (concentration of milk prior to cheese manqfacture). VI. 3.4.3 Summary ofh)llerfiltration membranes: asymmetric or composite thickness: sublayer,., 150 )lm; top layer,., 1 )lm pore size: <2nm driving force: pressure: brackish water 15 - 25 bar separation principle: membrane material: seawater 40 - 80 bar main applications: solution-diffusion cellulose triacetate, aromatic polyamide, poly(ether urea) (interfacial polymerisation) - desalination of brackish and seawater - production of ultrapure water (electronic industry) concentration offoodjuice and sugars (food industry), and the concentration of milk (dairy industry). VI . 3.5 Pressure retarded osmosis Pressure retarded osmosis (PRO) is a process derived from reverse osmosis. This process enables to generate energy from a concentration difference. The principle is shown in figure VI - 11. If a semipermeable membrane separates a concentrated salt solution from water or a dilute solution then osmosis occurs and water flows from the dilute solution (or pure water) to the concentrated solution. Only when a pressure is applied higher than the osmotic pressure water flows from the concentrated solution to the diluted solution. The osmotic water flow can be used to generate electricity by means of a turbine. The water flow at a pressure M> < ~1t can be described by VI - 21. (VI - 21) The power per unit membrane area is given by the product of flux and pressure difference, i.e. (VI - 28) The power is at maximum (W =Wmax) at dW/d(M> ) =0 ~ M> = 0.5 ~1t, which implies that Wmax = A4 1t2 (VI - 29) This equation clearly shows the effect of the osmotic pressure on the maximum power. This process was evaluated on the basis of experiments with existing membranes and seawater as saline solution. In this case about 1.5 W/m2 was produced but as a more
218 CHAPTER VI concentrated solution is used the energy production will increase drastically. However, there are a number of practical problems. fresh r---f';'.U·_-- swaaltienre water turbine Figure VI - 11 Principle of pressure retarded osmosis. - osmosis; because of osmosis the concentration of the concentrated solution will decrease and consequently the osmotic pressure decreases. - salt flux; when the membrane are not perfectly semipermeable (R < 100%) a salt flux occurs from the concentrated to the dilute side and as a result the osmotic pressure will decrease. - concentration polarisation. The severest problem is the occurrence of concentration polarisation (see chapter VII) which implies that the concentration at the membrane surface is different from that in the bulk (see figure VI - 12). The salt flux will cause an increased concentration in the sublayer, which can be considered as a stagnant layer, causing a decrease in effective osmotic pressure difference. This effect will decrease as Is => 0, which means that perfect semipermeable membranes (R = 100%) must be developed. concentration top laye~, sublayer 1 If ~ w Figure VI - 12. Concentration polarisation in pressure retarded osmosis.
MEMBRANE PROCESSES 219 VI. 3.5.1 Summary of pressure retarded osmosis- membranes: asymmetric or composite thickness: sublayer\"\" 150 Ilm; top layer\"\" Illm pore size: <2nm driving force: concentration difference (osmotic pressure) separation principle: solution-diffusion membrane material: cellulose triacetate, aromatic polyamide, main applications: poly(ether urea) (interfacial polymerisation) - production of energy VI . 3.6 Piezodialysis Another membrane process which uses pressure as the driving force is piezodialysis [5-7]. This process is applied with ionic solutes where in contrast to hyperfiltration, the ionic solutes permeate through the membrane rather than the solvent, which is usually water. A schematic drawing of the process is given in figure VI - 13. ee membrane circulating current Figure VI - 13. The transport of ions through a mosaic membrane during piezodialysis;iffi cation-exchange region; anion-exchange region So-called mosaic membranes must be used for this process. These are ion-exchange membranes possessing both cation-exchange and anion-exchange groups. Electroneutrality is maintained by the simultaneous passage of cations and anions through the membrane. Since ion transport is favoured relative to solvent transport, the salt concentration in the permeate is higher than that in the feed. This allows a dilute salt solution to be concentrated and a salt enrichment by a factor of two can be achieved. An increase in salt flux can be obtained by increasing the ion-exchange capacity of the membrane. Although the basic principle of the piezodialysis process has been demonstrated in the laboratory, it has not been employed on a commercial scale.
220 CHAPTER VI VI. 3.6.1 Summary of piezodialysis membranes: mosaic membranes (with cation-exchange regions thickness: adjacent to anion-exchange regions) pore size: \"\" few hundred !lm driving force: nonporous separation principle: pressure, up to 100 bar membrane material: ion transport (Coulomb attraction and electroneutrality) application: cation/anion-exchange membrane salt enrichment VI . 4 Concentration difference as the driving force VI . 4.1 Introduction In many processes, including those in nature, transport proceeds via diffusion rather than convection. Substances diffuse spontaneously from places with a high chemical potential to those where the chemical potential is lower. Processes which make use of a concentration difference as the driving force are gas separation, pervaporation, dialysis and liquid membrane processes (often it is better to speak of an activity difference rather than concentration difference). On the basis of differences in structure and functionality it is possible to distinguish between processes that use a synthetic solid (polymeric) membrane (gas separation, dialysis and pervaporation) and those that use a liquid (with or without a carrier) as the membrane. Whereas microfiltration, ultrafiltration and hyperfiltration are more or less similar processes, dialysis, gas separation and pervaporation differ quite considerably from each other. The basic feature that they have in common is the use of a nonporous membrane. It should be noticed that the term nonporous gives no information about the permeability. It was shown in chapter II that the permeability of a gas through an elastomeric and a glassy material may differ by more than five orders of magnitude, despite both materials being nonporous. This difference arises from large differences in segmental motion which is very restricted in the glassy state. The presence of crystallites can further reduce the mobility. A factor that enhances segmental mobility, or chain mobility in general, is the presence of low molecular penetrants. Increasing concentrations of penetrants (either gas or liquid) inside the polymeric membrane leads to an increase in the chain mobility and consequently to an increase in permeability (or diffusivity). The concentration of penetrant inside the polymeric membrane is determined mainly by the affinity between the penetrant and the polymer. In gas separation there is hardly any interaction between the gas molecules and the membrane material and the gas concentration in the membrane is very low. The gas molecules must diffuse through a rigid membrane structure with the state of the polymer being hardly effected by their presence. However, even for 'low affinity' penetrants of this type, there is a difference between the inert nitrogen and carbon dioxide, for example. In contrast, with liquid penetrants the solubility in the membrane may be appreciably higher which results in an enhanced chain mobility. An even greater interaction between liquid and membrane may occur in dialysis resulting in a much greater swelling of the polymer which allows relatively large molecules diffuse through this kind of open membrane. Figure VI - 14 shows schematically how the diffusion coefficient of a low molecular weight component changes as the degree of swelling of the membrane increases (the swelling of the membrane being defined as the weight fraction of penetrant inside the
MEMBRANE PROCESSES 221 membrane relative to the weight fraction of dry polymer). It will be seen that the diffusion coefficient can vary over the range 10-19 to 10-9 m2/s. This demonstrates quite clearly, that the mobility of the polymer chains increases with increasing swelling so that a situation is attained where the diffusivity is comparable to diffusion in a liquid (the diffusion coefficient in liquids is,., 10-9 m2/s). Thus swelling, as a result of interaction between the penetrant and the polymer, is a very important factor in transport through nonporous membranes. -12 10 -16 10 0.5 1.0 degree of swelling Figure VI - 14. Diffusivity as a function of the degree of swelling in nonporous polymers. Figure VI - 14 demonstrates that the diffusion coefficient can change by up to 10 orders of magnitude. Thus the diffusion coefficient of benzene in poly(vinyl alcohol) at zero penetrant concentration is less than 10-19 m2/s [8], whereas the diffusion coefficient of water in hydrogels is greater than 10-9 m2/s, which is virtually equal to value of the self- diffusion coefficient of water. VI . 4.2 Gas separation Gas separation is possible even with the two extreme types of membrane considered, i.e. porous and nonporous. The transport mechanisms through these two types of membrane, however, are completely different as discussed already in chapter V. VI . 4.2.1 Gas separation in porous membranes When gas transport takes place by viscous flow (as in the case of a microfiltration membrane, for example), no separation is achieved because the mean free path of the gas molecules is very small relative to the pore diameter. By decreasing the pore diameter of the pores in the membrane the mean free path of the gas molecules may become greater than the pore diameter. This kind of gas flow is called Knudsen flow which may be expressed by the equation:
222 CHAPTER VI J = 1t n r2 Dk ~p (VI - 30) RT t €. Mwwhere Dk , the Knudsen diffusion coefficient, is given by Dk = 0.66 r ~ 8 RMwT 1t T and are the temperature and molecular weight, respectively and r is the pore radius. Eq. VI - 30 shows that the flow is inversely proportional to the square root of the molecular weight and the latter is the only parameter which determines the flow for a given membrane and a given pressure difference. Hence, the separation of two gases by a Knudsen flow mechanism depends on the ratio of the square root of their corresponding molecular weights. This means that low separation factors are generally obtained. High separation can only be achieved via a cascade operation involving a number of modules connected together (see chapter VIII). For economical reasons this is very unattractive and thus the only commercial application of this method to date has been the enrichment of uranium hexafluoride (235UF6 ), a very expensive material. The separation factor obtained in the separation of 235UF6 from 238UF6 is extremely low (the ideal separation factor is 1.0043, but this factor will not be attained in the practical situation). A plant employing this application method using porous ceramic membranes operates in France (at Tricastin). However, there is another aspect to Knudsen flow. Where the transport of gases occurs through nonporous membranes, as will be discussed in the following section, Knudsen flow is not involved. However, when these nonporous membranes are used in a composite membrane where a dense top layer is supported by a porous substructure, Knudsen flow may contribute to the total flow depending on the pore sizes in the sublayer. VI . 4.2.2 Gas separation throu~h nonporous membranes Gas separation through nonporous membranes depends on differences in the permeabiIities of various gases through a given membrane. Fick's law is the simplest description of gaS' diffusion through a nonporous structure, i.e. J = _Ddc (VI - 31) dx where J is the flow rate through the membrane, D is the diffusion coefficient and the driving force dc/dx is the concentration gradient across the membrane. Under steady-state conditions this equation can be integrated to give: J = D (co - cd (VI - 32) €. where Co and Ct are the concentrations in the membrane on the upstream side and downstream side, respectively. The quantity€. is the thickness of the membrane. The concentrations are related to the partial pressures by Henry's law which states that a linear relationship exists between the concentration inside the membrane and the (partial) pressure of gas outside the membrane, i.e. c = S. P (VI - 33)
MEMBRANE PROCESSES 223 where S is the solubility coefficient. Henry's law is mainly applicable to amorphous elastomeric polymers for the solubility behaviour is very often much more complex below the glass transition temperature, as has been described in chapter V. Combining eq. VI - 32 with eq. VI - 33 gives eJ = D S (Po - pt) (VI - 34) an equation which is generally used for the description of gas permeation through membranes. The product of the diffusion coefficient D and the solubility coefficient S is called the permeability coefficient P, i.e. P=D. S (VI - 35) so that eq.VI - 34 can be written as: (VI - 36) eJ = p (Po - Pt) Eq. VI - 36 shows that the flow rate across a membrane is proportional to the difference in (partial) pressure and inversely proportional to the membrane thickness. The ideal selectivity is given by the ratio of the permeability coefficients: (Xi/j ideal Pi (VI - 37) p. J With a number of gaseous mixtures, the real separation factor is not equal to the ideal separation factor because of plasticisation which may occur at high (partial) pressures when a permeating gas exhibits a high chemical affinity for the polymer. Because of such plasticisation, the permeability increases but the selectivity decreases generally. feed ~I--~.c•ompressor ~ -- --1- --:~I--r::~nt.a.t.e permeate vacuum pump Figure VI - 15. Schematic drawing of a gas separation process. The real separation factor also depends on the pressure ratio across the membrane. In the case of a high pressure ratio the separation efficiency is a maximum when the pressure ratio is high (or Pe/po ~ 0) and the selectivity decreases as the pressure ratio decreases (see also chapter VIII). The driving force can be established either by applying a high pressure on the feed side and/or maintaining a low pressure on the permeate side. A schematic drawing of such a gas separation process is shown in figure VI - 15.
224 CHAPTER VI VI . 4.2.3. Aspects of sIWaration The permeability coefficient P is a very characteristic parameter which is often described as an intrinsic parameter easily available from simple permeation experiments with membranes of known thickness (using eq. VI - 34 ). The permeability coefficient is often given in Barrer units. (lBarrer=lO- lO cm3(STP).cm.cm-2.s-1.cmHg-1 = 0.76 10- 17 m3(STP).m.m-2.s·1.Pa-1). The dimensions for permeability coefficients indicate that they depend on the membrane thickness, the membrane area and the driving force. However, in those cases where Henry's law does not apply the permeability coefficient P is no longer a constant but is related to the driving force, i.e. varying the pressure leads to different values for P. Nevertheless, the permeability coefficient is a convenient parameter for comparing the efficiency of different membranes for a given separation and for comparing the behaviour of various gases in a given membrane. To describe the fundamentals of gas separation, however, other factors relating to the nature of the polymer (i.e. chemical structure) need to be considered. Two parameters are important in this context: i) the glass transition temperature and ii) the crystallinity. The glass transition temperature determines whether a polymer is in the glassy or in the rubbery state. Segmental motion is limited for an amorphous polymer in the glassy state, whereas in the rubbery state enough thermal energy is available to allow rotation in the main chain. The glass transition temperature is mainly determined by chain flexibility TABLE VI - 7. The permeability of carbon dioxide and methane in various polymers [53-55] polymer Pe02 (Barrer) polytrimethylsilylpropyne 33100 2.0 silicone rubber 3200 3.4 4.6 natural rubber 130 8.5 11.2 polystyrene 11 15.1 polyamide (Nylon 6) 0.16 26.7 28.0 poly(vinyl chloride) 0.16 32.0 31.6 polycarbonate (Lexan) 10.0 31.0 polsulfone 4.4 45.0 poly(ether sulfone) (Victrex) 7.4 64.0 polyethyleneterephthalate (Mylar) 0.14 cellulose acetate 6.0 poly(ether imide) (Ultem) 1.5 polyimide (Kapton) 0.2 1 Barrer =10-10 cm3(STP).cm.cm-2.s-1.cmHg-l and chain interaction. These parameters have been discussed in detail in chapter II. In general, permeability through a rubbery material (elastomer) is much higher relative to glassy polymers because of the higher mobility of the chain segments. In contrast, the selectivity of glassy polymers is higher. Table VI - 7 lists the permeability of carbon dioxide (in Barrer) and the ratio of the permeabilities (the ideal selectivity) of carbon dioxide and methane in various polymers.
MEMBRANE PROCESSES 225 TABLE VI - 8. The permeability of oxygen and nitrogen for some elastomers and glassy polymers [53-55] Polymer Tg P02 PNZ U ideal (CC) (Barrer) (Barrer) (POiPNZ) PPO 210 16.8 3.8 4.4 10040.0 6745.0 PTMSP =200 1.5 11.2 3.3 3.4 ethylcellulose 43 37.2 8.9 4.2 0.3 polymethylpentene 29 1.6 1.2 5.4 4.0 1.0 3.3 polypropylene -10 2.9 0.14 2.9 0.4 2.9 polychloroprene -73 polyethylene LD -73 polyethylene HD -23 The results listed in table VI - 7 indicate that elastomers exhibit high permeabilities and low selectivities whereas glassy polymers show much lower permeabilities but generally higher selectivities. There is however no unique relationship between the glass transition temperature and permeability, or in other words the permeability of a rubber is not 'a priori' greater than that of a glassy polymer. Table VI - 8 summarises some examples where the permeability of glassy polymers are higher than those of elastomers. These examples have only been given in order to demonstrate that exceptions exist to the rule that the permeability of elastomers is higher than that of glassy polymers. The rule applies in general of course, with the vast majority of elastomers having a higher permeability than most of the glassy polymers. Only in those cases where the fractional free volume of the polymer is high (e.g. polytrimethylsilylpropyne) are high permeabilities obtained. The basic concept of gas separation is governed by the permeability coefficient (P) which is equal to the product of the solubility (S) and the diffusivity (0). In comparison to liquids the affinity of gas molecules towards a polymer is generally much lower and hence the solubility of gases in polymers is quite low (generally < 0.2%). The solubility is mainly determined by the ease of condensation. Because larger molecules condense more readily, their solubility increases. This can be illustrated by the example of the noble gases. Such gases show no polymer interaction and their solubility is determined only by their ease of condensation. Hence the solubility increases with increasing size of the gas molecules (and with increasing critical temperature or boiling temperature) in the sequence: neon, argon, krypton, and xenon [9]. Thus the solubility of neon in silicone rubber is 0.04 cm3(STP).cm-3.atm-1 whereas for krypton a value of 1.0 cm3(STP).cm-3.atm-1 is found [9]. Solubility of a given gas molecule increases as its polymer affinity increases. For example the solubility of carbon dioxide in hydrophilic polymers is generally higher than in more hydrophobic polymers. The other factor affecting permeability is the diffusivity. It depends mainly on two factors: the molecular size of the gaseous penetrant and the choice of the polymer. The size of the gas molecule is reflected in the diffusion coefficient, i.e. the smaller its size the higher the diffusion coefficient. Indeed, a close examination of the dimensions of gas molecules provides some interesting results. Table VI - 9 summarises the kinetic diameters of some relevant gas molecules [13].
226 CHAPTER VI TABLE VI - 9. The kinetic diameter of some gas molecules [13] gas diameter molecule (A) He 2.6 Ne 2.75 H2 2.89 NO CO2 3.17 C2H2 3.3 Ar 3.3 °2 N2 3.4 CO 3.46 CH4 C2H4 3.64 C 3Hg 3.76 3.80 3.9 4.3 Thus, although the molecular weight of oxygen is greater than that of nitrogen, the molecular dimensions of oxygen are smaller. Hence when the permeability is considered in terms of diffusivities oxygen will generally have a higher permeability than nitrogen. The tables of permeability coefficients (tables VI - 7 and VI - 8) demonstrate that this is indeed the case, not only for glassy polymers but also for elastomers. Only in glassy polymers is the separation factor generally higher. It has already been shown in chapter V that the thermodynamic diffusion coefficient can be expressed as: Dr = kfT (VI - 38) where f is the frictional coefficient. Stokes' law demonstrates that the frictional coefficient is related to the size of the diffusing molecule by: f = 61tllr (VI - 39) Combination of eq. VI - 38 with eq. VI - 39 for ideal systems (DT =D) gives (VI - 40) This relationship shows that the diffusion coefficient is inversely proportional to the molecular size. Although not very accurate for the diffusion of gases in polymers, this relationship does illustrate the link between the diffusion coefficient and the size. Relative small differences in size have a very large effect on the diffusion coefficient. For example, the diffusion coefficient of neon (Mw 20) in polymethylmethacrylate (PMMA) is approximately 10-10 m2s-1 and for krypton (Mw 83.8) approximately 10-12 m2s-1 [9].
MEMBRANE PROCESSES 227 The diffusion coefficient also depends strongly on the nature of the polymer. For example the diffusion coefficient of krypton in polydimethylsiloxane is about 10-9 m2s-1 while for the same gas in PVA values of 1013 m2s-1 have been reported, i.e. four orders of magnitude lower [9]. A comparison of the separation properties requires an evaluation not of the solubilities and diffusivities but rather their respective ratio. Table VI - 10 lists the ratios of the solubilities (S), diffusivities (D) and permeabilities (P) for CO2 and CH4 in some glassy polymers [14]. TABLE VI - 10. Ratios of the diffusivities, solubilities and permeabilities of CO2 and CH4 in various polymers [14] polymer PCOzIPCH4 cellulose acetate 4.2 7.3 30.8 polyimide 15.4 4.1 63.6 polycarbonate 6.8 3.6 24.4 polysulfone 8.9 3.2 28.3 The affinity of carbon dioxide for a given polymer is (much) higher than that of methane. This can be clearly seen from table VI - 10, where in cellulose acetate or other ester- containing polymers the solubility of CO2 is especially high and a high solubility ratio can be found. However, it appears that high selectivities are not necessarily based on large differences in solubility, but that diffusivity or changes in diffusivity in particular have a much stronger effect on the selectivity. Thus, the polyimide shown in the table (Kapton) is a glassy polymer with a very rigid structure. Table VI - 10 shows that for such polymers it is mainly the diffusivity ratio which determines the selective transport, suggesting the existence of a microstructures which is able to discriminate on a molecular level. Because molecules of almost the same size can be separated this implies that openings (in terms of free volume) with very definite dimensions exist within the polymeric matrix which allow smaller molecules to pass (much) more readily than larger ones. These kinds of very rigid structure are quite similar to those in zeolites (or molecular sieves) which also contain very definite structures. Such behaviour may be observed not only for the CO2/CH4 separation but also for the separation of oxygen and nitrogen. Almost all polymers have selectivity factors (or PoJPN2) between 2 and 6 [15], but some rigid glassy polymers, similar to the polyimide mentioned above, have higher selectivities. It is assumed that it is the very definite pore structure which exclude the larger nitrogen molecule to a greater extent than the smaller oxygen molecule. Hence, separation is determined by the selective diffusion of oxygen to nitrogen rather than specific interaction. This means that highly selective polymers capable of separating permanent gases should be glassy polymers rather than elastomers. Furthermore, the microstructure seems to be much more important than the existence of specific interactions. However, the permeability is often very low and differences in permeability can be as much as six orders of magnitude (compare the permeabilities of various gases through poly(vinyl alcohol) or polyacrylonitrile with that through polydimethylsiloxane). Up to this point it has been demonstrated that the permeability of a gas depends very
228 CHAPTER VI much on the choice of the polymer. However, when different gases are used with the same polymer (membrane) large differences in permeability can be observed. This is especially true for organic vapours where differences can extend over six orders of magnitude. The (ocedifference between a gas and a vapour lies in the fact that vapours are condensable under standard conditions and 1 bar). Table VI - 11 gives the permeabilities of various agtasaensaacntidvvitaypoofuars=in1p(pol=ydpi~m.ethylsiloxane [16] , the vapour values having been measured TABLE VI - 11. Permeabilities of various gases and vapours in polydimethylsiloxane [16]. Component Permeability (Barrer) nitrogen 280 oxygen 600 methane 940 carbon dioxide 3200 45,000 ethanol 168,000 methylene chloride 284,000 chloroform 200,000 carbon tetrachloride 248,000 1,2-dichloroethane 247,000 1,1, I-trichloroethane 614,000 trichloroethylene 1,460,000 toluene Although the kinetic dimensions of the various organic vapour molecules are much larger than those of oxygen and nitrogen, the permeabilities are much higher. Because the permeability is determined by the solubility and the diffusivity, this suggests that the high permeability originates from a much higher solubility. Organic vapour molecules exert a plasticising action on the polymer, i.e. the polymer chains become much more flexible, alternatively the free volume increases considerably. This effect increases with increasing solubility and an exponential relation is often found. It is also possible to derive an exponential relationship from the free volume theory as was described in chapter V. The following empirical relationship is often used: (VI - 41) where Do is the diffusion coefficient at zero penetrant concentration, y is a constant related to the plasticising effect of the penetrant on the polymer and <1> is the volume fraction of the penetrant in the membrane. The concentration dependence of the diffusion coefficient is not the same for all polymers. Do is mainly determined by the penetrant size and shape, and by the choice of the polymer. For a given polymer, Do decreases with increasing penetrant size: for example, the value of Do for methanol in poly(vinyl alcohol) is about three orders of magnitude larger than the Do value for n-propanol [9]. For a given penetrant, Do increases with increasing chain flexibility. Hence Do increases drastically in going from a glassy polymer to an elastomer; for example, the Do
MEMBRANE PROCESSES 229 value of benzene in poly(vinyl alcohol) is about ten orders of magnitude lower than that of benzene in polydimethylsiloxane (silicone rubber). A correlation between Do and Tg has been proposed but glassy polymers as poly(dimethylphenylene oxide) and polytrimethylsilylpropyne also show very high permeabilities which stem from very high diffusivi ties. The solubility of organic vapours in (glassy) polymers is generally much higher than that of permanent gases in the same polymer. Whereas Henry's law can be used in the latter case, with organic vapours the solubility can be described by Flory-Huggins thermodynamics, as for liquids. VI . 4.2.4 Membranes for gas separation Table VI - 7 indicates that the permeability of a given gas molecule in various polymers can change by more than six orders of magnitude. Equally, table VI - 11 also shows that for a given polymer the permeability of various gas and vapour molecules can change over six orders of magnitude. This large variation in permeability shows that in principle many materials can be used as a membrane. Gas separation is not only based on permeability but also on the selectivity, which is equal to the ratio of the permeabilities for gas mixtures. For separation problems involving large differences in interaction, e.g. gases from vapours, the permeability ratio is usually large (see table VI - 11) and for this reason a highly permeable material may be chosen. In general, these are elastomers such as silicone rubber or natural rubber. Elastomers show rather low selectivities for some separations and glassy polymers with a much lower permeability are often used. It can be seen from eq. VI - 29 that the permeation rate (= P / t) varies inversely with membrane thickness. For this reason the permeation properties can be optimised by minimising the effective membrane thickness. Therefore two types of membranes are very suitable for gas separation: - asymmetric membranes - composite membranes Asymmetric membranes are mainly prepared by immersion precipitation whereas this technique is also used for the sublayer in composite membrane upon which a very thin selective layer is deposited by one of the following techniques: - dip-coating - interfacial polymerisation - plasma polymerisation All these techniques have been described in chapter III. In both asymmetric and composite membranes the hydrodynamic resistance is determined largely by the thin dense top layer. This top layer must be absolutely defect-free, since a few defects can significantly reduce the selectivity without having much influence on the flux. In addition, the following requirements are necessary for the porous support layer: - it must provide mechanical support for the top layer - it must have an open porous network to minimise resistance to mass transfer (no closed pores !) - it must not contain macrovoids (weak spots for high-pressure applications) It is very difficult to make a defect-free thin top layer from a glassy polymer. However, two phase inversion methods can be used to prepare a defect-free asymmetric membrane, i.e. the dual bath method [17] and the evaporation method [18,19]. Another elegant method of preparing a defect-free 'asymmetric' membrane is to deposit a coating of a highly permeable polymer upon an asymmetric membrane containing some defects. This
230 CHAPTER VI coating layer plugs the surrace pores resulting in a membrane without defects [20]. It is also possible to reduce the top layer thickness further to increase the permeation rate. At the same time it is interesting to know how many imperrections can be allowed without losing too much in selectivity. The effectiveness of this procedure can be easily be demonstrated by considering a resistance model [20]. Figure VI - 16 shows a schematic representation of an asymmetric membrane and the corresponding electrical circuit analogue. toplayer pore (3) (2) Figure VI - 16. Schematic representation of an asymmetric membrane and the corresponding electrical circuit analogue. It is obvious that the surface porosity must be negligible otherwise the selectivity will decrease dramatically. By applying a thin coating layer upon these asymmetric membrane these defects will be plugged. Although an extra resistance has now been introduced. the resistance of the plugged pores is much higher than of the open pores so that a much better perrormance results. This can also be demonstrated by means of a resistance model which can explain the effectiveness of a highly permeable low-selective defect-free coating layer. The gas flow through a membrane per unit area per unit time is given by eJ = £ ~p or (VI - 42) The overall permeability P can be expressed in terms of resistances as: (VI - 43) where Rtot is the total membrane resistance. For the uncoated membrane the total resistance Rtot.un is given by (see figure VI - 16)
MEMBRANE PROCESSES 231 (1 ) porous substructure .-----.. (4) toplayer subs tructu re (2) plugg pore (3) Figure VI - 17. Schematic representation of a coated asymmetric membrane and the corresponding electrical circuit analogue. (VI - 44) while for the coated layer the total resistance Rtot,e is given by (see figure VI - 17) (VI - 45) If it is assumed that the resistance in the sublayer ( R4 ) is negligible, the flux of the uncoated (Jun ) and coated membranes (Je ) may be written as: Jun P2 + PI (A2 / A3 ) (VI - 46) t.p ~ and (VI - 47) Here, tl is the thickness of the coating layer, t2 is the thickness of the top layer in the substructure, A2 is the total pore area and A3 is the surface area of the (solid) polymer (the ratio AJA3 gives the surface porosity). Figure VI - 18 gives the flux and selectivity as a function of the surface porosity for the uncoated and coated membranes. From this figure it can be seen that a coating procedure with a very permeable low-selective polymer is very effective in obtaining a defect-free layer. For the uncoated membrane any defect leads to a decrease in selectivity, whereas in coated membranes surface porosities (defects) up to 10-4 may be allowed without any decrease in selectivity. The permeability is hardly affected by the presence of
232 CHAPTER VI (Pit)rei a coated 102 10-8 10-6 10-8 10-6 10-4 surface porosity surface porosity Figure VI - 18. Selectivity and flux as a function of the surface porosity for coated and uncoated membranes on the basis of the resistance model. the coating layer. Hence this method enables the uncoated membrane to be optimised with respect to flux. Although some defects might be present because of the reduction in the top layer thickness, these will have no effect because of the coating layer. It should be emphasised again that the performance of these composite membranes is determined by the asymmetric membrane (or the intrinsic properties of the polymer used to prepare this membrane) and the only function of the coating layer is to plug the pores (defects). There is also another type of composite membrane consisting of a sublayer and top layer too, but here transport through the thin top layer is the rate-determining step. In terms of the resistance model this means that the resistance of the thin top layer is much greater than that of the sublayer and that the separation performance is determined by the intrinsic properties. Sometimes, a highly permeable third layer, e.g., polydimethylsiloxane, is used between the sublayer and top layer and serves as an intermediate layer or 'gutter'. When the surface of the sublayer is highly porous, it is often difficult to deposit a thin selective coating directly. Also when the top layer is composed of a glassy polymer it is often difficult to obtain this layer defect-free. Under these circumstances the three-layer membrane or 'double composite' membrane may be a good approach [21]. Several methods capable of depositing the thin selective layer upon the support have been described already, i.e. dip-coating and plasma polymerisation. Plasma polymerisation produces very thin top layers whose structures are difficult to control whereas the thickness can easily be varied in a dip-coating procedure by varying the polymer concentration in the coating solution. With plasma polymerisation it is even necessary to have an intermediate nonporous layer to obtain a defect-free membrane. VI . 4.2.5 Applications Multi-stage gas separation is not very attractive in general (see chapter VIII) for ordinary gas separation mixtures. The membranes need to combine a high flux with a reasonably high selectivity. High fluxes or high permeabilities are generally related to low selectivities and in order to discuss gas separation applications a classification into high permeable and low permeable materials will be made:
MEMBRANE PROCESSES 233 - High permeable materials are used if high selectivities are not required, as for example the production of oxygen enriched air for medical applications, combustion processes, and sterile air for aerobic fermentation processes. Another application is the separation of organic vapours from non-condensable gases such as nitrogen (air!), where high selectivities may be obtained with highly permeable materials. The permeability of any material for nitrogen is much lower than that for any organic vapour and hence it is of advantage to select a high permeable material for this application. - If reasonable selectivity is required then then low permeable materials based on glassy polymers will be employed. In practice a balance must be struck between permeability and s*eleCcOtivyiCtyH. A4 large number of applications can be mentioned. This kind of separation problem arises in many applications: the purification of CH4 from landfill drainage gas, the purification of CH4 from natural gas and the recovery of CO2 in e*nhHa2ncoerdHoeilfrreocmovoetrhye.r gases Hydrogen and helium have relatively small molecular sizes compared to other gases and exhibit high selectivity ratios in glassy polymers. Applications can be found in the recovery of H2 from purge gas streams in ammonia synthesis, petroleum refineries and m* eHth2aSn/oCl Hsy4nthesis. Besides CO2 in natural gas H2S is often present in appreciable concentrations. The concentration of this very toxic, highly corrosive gas has to be reduced to less than 0.2%. * ~/N2 Separation can be effected to obtain both oxygen-enriched air and nitrogen-enriched air. Nitrogen-enriched air (95 - 99.9%) can be used as an inert gas in the blanketing of fuel t*anHks2.0 from gases A* llSk0in2dfsroomf gsamseoskceagnabse dried by the removal of water. Desulfurisation of smoke gas VI. 4.2.6 Summary of gas separation membranes: composite or asymmetric membranes with an thickness: elastomeric or glassy polymeric top layer pore size: '\" 0.1 to few Ilm (for top layer) driving force: nonporous (or porous < 11lm) separation principle: pressure, upstream to 100 bar or vacuum downstream membrane material: solution/diffusion (nonporous membranes) application: Knudsen flow (porous membranes) elastomer: polydimethylsiloxane, polymethylpentene glassy polymer: polyimide, polysulfone - H2 or He recovery - CHJC02, H2S - °2/N2 - removal of H20 (drying) - organic vapours from air
234 CHAPTER VI VI . 4.3 Pervaporation Pervaporation is a membrane process in which a liquid is maintained at atmospheric pressure on the feed or upstream side of the membrane and where the permeate is removed as a vapour because of a low vapour pressure existing on the permeate or downstream side. This low (partial) vapour pressure can be achieved by employing a carrier gas or using a vacuum pump. The (partial) downstream pressure must be lower than the saturation pressure at least. A schematic drawing of this process is shown in figure VI - 19. -fee-d ~.~r----~ret~ent~ate~ feed retentate penneate condenser ~camer gas permeate Figure VI -19. Schematic drawing of the pervaporation process with a downstream vacuum or an inert carrier-gas. Essentially, the pervaporation process involves a sequence of three steps: selective sorption into the membrane on the feed side selective diffusion through the membrane desorption into a vapour phase on the permeate side Pervaporation is a complex process in which both mass and heat transfer occurs. The membrane acts as a barrier layer between a liquid and a vapour phase implying that a phase transition occurs in going from the feed to the permeate. At the very least this means that the heat of vaporisation must be supplied. Because of the existence of a liquid and a vapour pervaporation is often considered as a kind of extractive distillation process with the membrane acting as a third component. However, the separation principle in distillation is based on the vapour-liquid equilibrium whereas separation in pervaporation is based on differences in solubility and diffusivity. The vapour-liquid eqUilibrium influences the separation characteristics because it directly affects the driving force. Figure VI - 20 compares distillation (vapour-liquid equilibrium) with pervaporation for an ethanol-water mixture at 20ce. The pervaporation experiments were carried using a polyacrylonitrile membrane. The description of transport given above has already indicated that the desorption step hardly contributes to transport resistance. Transport can be described by means of a solution-diffusion mechanism where the selectivity is determined by selective sorption and/or selective diffusion. In fact, the same type of membrane or membrane material can be
MEMBRANE PROCESSES 235 weight fraction 0.5 of water in vapour Ipervaporation / / 0.5 / / / vapour-liquid 0.5 equilibrium / / 0.5 weight fraction of water in liquid Figure VI - 20. Distillation (vapour-liquid equilibrium) and pervaporation characteristics for an ethanol-water mixture at 20OC. Pervaporation was carried out using a polyacrylonitrile membrane [56]. used for both gas separation or pervaporation. However, the affinity of a liquid towards a polymer is generally much higher than that of a gas in a polymer so that the solubility is much higher. This effect could already be noticed in the case of organic vapours which exhibit much higher permeabilities than permanent gases such as nitrogen. In gas separation, the selectivity towards a mixture can be estimated from the ratio of the permeability coefficients of the pure gases. However, with liquid mixtures the separation characteristics are far different from those of a pure liquid because of coupling phenomena and thermodynamic interactions. The low solubility of gases in polymeric materials (at T < Tg) can be described by Henry's law. The much higher solubility of liquids implies that Henry's law is no longer obeyed, and the Flory-Huggins theory is commonly used to provide an adequate description of the solubility of liquid mixtures and pure liquids into a polymeric material. The permeability of a given component i from a mixture of components i and jean be expressed as a function of the diffusivity (D) and the solubility (S). With liquids the main difference from gases is that the diffusivity and the solubility are not constants but are strongly dependent on the feed composition; (VI - 48) If another component k is taken instead of component j, both the diffusivity Dj and the solubility Sj are changed. In addition, this clearly indicates that the permeability coefficient is not a constant but strongly dependent on the composition of the liquid mixture. The following (extreme) examples demonstrate this. If poly(vinyl alcohol) is used for the separation of ethanol-
236 CHAPTER VI water mixtures, two compositions can be distinguished: a low water concentration and a low alcohol concentration. With this low alcohol concentration (say less than 10%) the membrane is highly swollen and hardly any selectivity is obtained. With the low water concentration (say less than 10%), this same polymer membrane shows a high selectivity towards water and exhibits a reasonable flux. Another example is that of a mixture which consists of two components which are not miscible with each other over the whole composition range, e.g. trichloroethylene- water. Pervaporation can be used to remove a small amount of water from trichloroethylene or to remove small amounts of trichloroethylene from water. If silicone rubber (polydimethylsiloxane) is used as a membrane, good results are obtained as far as the removal of small amounts of trichloroethylene from water is concerned, showing that high selectivities exist for trichloroethylene combined with reasonable permeation rates. When the same membrane material is used to remove water from almost pure trichloroethylene, the membrane becomes highly swollen and cannot continued to be used. Thus in order to remove traces of water another material has to be chosen, e.g. poly(vinyl alcohol). These extreme examples indicate the influence of composition on the membrane performance. VI . 4.3.1. Aspects of separation For single component transport, simple transport equations can be derived from linear flux- force relationships: J =- L· dd/-xli (VI - 49) 1 1 where Li is a proportionality or phenomenological coefficient. The chemical potential is given by (VI - 50) with (VI - 51) pi being the saturation pressure of component i and Pi is its vapour pressure. Because d/-li = RT dPi (VI - 52) dx Pi dx (VI - 53) eq. VI - 49 now becomes J - -LiPR-iT- ddpxi 1 - eTaking differences instead of differentials (dP/dx \"\" ~Pi /~) where ~ is the membrane thickness and Pi = (Li . RT)/Pi ,eq. VI - 53 becomes
MEMBRANE PROCESSES 237 (VI - 54) This liquid transport equation is the same as that for gas transport (see eq. VI - 30), the only difference is that the value of the permeability coefficient is not the same. Furthermore, for liquid mixtures the ratio of the permeability coefficients is not equal to the selectivity factor. Eq. VI - 54 illustrates the important parameters involved, the permeability dcoeteafiflicbieenlotwb.eiOngthaermpeamrabmraentee-rsoor fmiantteerrieaslt-baarseedthpeaeraffmecettievrewmhiecmhbwrailnlebetheicxkpnlaeisnseed. in more and the partial pressure difference ~p. The permeation rate is inversely proportional to the membrane thickness and proportional to the partial pressure difference across the membrane. The vapour pressure on the permeate side is minimal (Pi,2 => 0) when a vacuum is applied in combination with liquid nitrogen temperatures (- 196CC or 77 K). In this case the driving force is determined completely by the vapour pressure of the feed liquid, which in tum is determined by the temperature of the feed. Liquid nitrogen is far from attractive for commercial applications, but the vapour can also be condensed with cooling water at about room temperature. However, under these circumstances it is necessary to increase the temperature of the liquid feed in order to maintain sufficiently large driving force. Eq. VI - 54 can easily be rewritten in terms of a diffusion coefficient. Combining eqs. VI - 49 and VI - 50, the following equation is obtained J1· = - L· RT dldnxai (VI - 55) 1 or (VI - 56) The activity ~ of a component in a polymeric membrane can be described by Flory- Huggins thermodynamics [24]. Thus, the activity of a component (index i) in a polymer (index j) is given by In a·1 = In '\"1''1. + (I - Y1N·J)' '\"I''.J + X\"1J \"'+''.J2 (VI - 57) where <Pi is the volume fraction of the liquid inside the polymer, <P j is the volume fraction of polymer and Xij is the Flory-Huggins interaction parameter. For an ideal system (Vi = Vj and Xij = 0), differentiation of eq. VI - 57 with respect to <P i gives dlnai l d~ (VI - 58) If we define the concentration dependent diffusion coefficient Di (c) as
238 CHAPTER VI ~ (C) = Li R T dlnai (VI - 59) dx then writing concentrations c instead of volume fractions <l> and combining eqs. VI - 56, VI - 58 and VI - 59 gives Ji = - Di (c) ~dx (VI - 60) In eq. VI - 60, Di (c) is the diffusion coefficient of component i in the polymer fixed frame of reference and is a function of the concentration. The liquid generally swells the polymer to a certain extent during pervaporation. Such swelling is anisotropic, since the the liquid concentration on the feed side of the membrane is a maximum whereas on the permeate side the swelling is almost zero. Figure VI - 21 gives a schematic drawing of the concentration profile, or in this case an activity profile. It is assumed that thermodynamic equilibrium exists at the interfaces, i.e. the activity of the liquid in the feed and in the membrane are the same (for pure liquids this means that the activity is unity). When the vapour pressure on the permeate side is very low (or P2/po ~ 0), the activity or concentration varies quite considerably over the membrane and the driving force is a feed membrane penneate (liquid) (vapour) 1.0 tactivity Figure VI - 21. Activity profile of a pure liquid across a membrane. maximum. Consequently, the concentration-dependent diffusion coefficient will also change quite considerably across the membrane. Indeed, an exponential relationship is often used to express the concentration dependence of the diffusion coefficient, i.e. (VI - 61) where Do,i is the diffusion coefficient at c=>O and 'Y is a plasticising constant expressing the plasticising action of the liquid on segmental motion. Combining eqs. VI - 60 and VI - 61 and integrating across the membrane using the boundary conditions Ci = ci, 1m at x = 0 eci = 0 at x =
Search
Read the Text Version
- 1
- 2
- 3
- 4
- 5
- 6
- 7
- 8
- 9
- 10
- 11
- 12
- 13
- 14
- 15
- 16
- 17
- 18
- 19
- 20
- 21
- 22
- 23
- 24
- 25
- 26
- 27
- 28
- 29
- 30
- 31
- 32
- 33
- 34
- 35
- 36
- 37
- 38
- 39
- 40
- 41
- 42
- 43
- 44
- 45
- 46
- 47
- 48
- 49
- 50
- 51
- 52
- 53
- 54
- 55
- 56
- 57
- 58
- 59
- 60
- 61
- 62
- 63
- 64
- 65
- 66
- 67
- 68
- 69
- 70
- 71
- 72
- 73
- 74
- 75
- 76
- 77
- 78
- 79
- 80
- 81
- 82
- 83
- 84
- 85
- 86
- 87
- 88
- 89
- 90
- 91
- 92
- 93
- 94
- 95
- 96
- 97
- 98
- 99
- 100
- 101
- 102
- 103
- 104
- 105
- 106
- 107
- 108
- 109
- 110
- 111
- 112
- 113
- 114
- 115
- 116
- 117
- 118
- 119
- 120
- 121
- 122
- 123
- 124
- 125
- 126
- 127
- 128
- 129
- 130
- 131
- 132
- 133
- 134
- 135
- 136
- 137
- 138
- 139
- 140
- 141
- 142
- 143
- 144
- 145
- 146
- 147
- 148
- 149
- 150
- 151
- 152
- 153
- 154
- 155
- 156
- 157
- 158
- 159
- 160
- 161
- 162
- 163
- 164
- 165
- 166
- 167
- 168
- 169
- 170
- 171
- 172
- 173
- 174
- 175
- 176
- 177
- 178
- 179
- 180
- 181
- 182
- 183
- 184
- 185
- 186
- 187
- 188
- 189
- 190
- 191
- 192
- 193
- 194
- 195
- 196
- 197
- 198
- 199
- 200
- 201
- 202
- 203
- 204
- 205
- 206
- 207
- 208
- 209
- 210
- 211
- 212
- 213
- 214
- 215
- 216
- 217
- 218
- 219
- 220
- 221
- 222
- 223
- 224
- 225
- 226
- 227
- 228
- 229
- 230
- 231
- 232
- 233
- 234
- 235
- 236
- 237
- 238
- 239
- 240
- 241
- 242
- 243
- 244
- 245
- 246
- 247
- 248
- 249
- 250
- 251
- 252
- 253
- 254
- 255
- 256
- 257
- 258
- 259
- 260
- 261
- 262
- 263
- 264
- 265
- 266
- 267
- 268
- 269
- 270
- 271
- 272
- 273
- 274
- 275
- 276
- 277
- 278
- 279
- 280
- 281
- 282
- 283
- 284
- 285
- 286
- 287
- 288
- 289
- 290
- 291
- 292
- 293
- 294
- 295
- 296
- 297
- 298
- 299
- 300
- 301
- 302
- 303
- 304
- 305
- 306
- 307
- 308
- 309
- 310
- 311
- 312
- 313
- 314
- 315
- 316
- 317
- 318
- 319
- 320
- 321
- 322
- 323
- 324
- 325
- 326
- 327
- 328
- 329
- 330
- 331
- 332
- 333
- 334
- 335
- 336
- 337
- 338
- 339
- 340
- 341
- 342
- 343
- 344
- 345
- 346
- 347
- 348
- 349
- 350
- 351
- 352
- 353
- 354
- 355
- 356
- 357
- 358
- 359
- 360
- 361
- 362
- 363
- 364
- 365
- 366
- 367
- 368
- 369
- 370
- 371
- 372
- 373
- 374