<<

2508 Macromolecules 2003, 36, 2508-2515

Phase Behavior Prediction of Ternary Polymer

Juan A. Gonzalez-Leon and Anne M. Mayes* Department of and Engineering, Massachusetts Institute of Technology, 77 Massachusetts Ave., Cambridge, Massachusetts 02139 Received June 24, 2002; Revised Manuscript Received February 3, 2003

ABSTRACT: A compressible regular free energy model for multicomponent polymer blends is developed and used to obtain spinodal curves for three component systems, employing only pure component properties such as coefficients of thermal expansion, , and parameters. The ability to predict, at least qualitatively, the behavior of the ternary polymer mixtures tetramethylpolycar- bonate (TMPC)/polycarbonate (PC)/poly(styrene-r-acrylonitrile) (SAN), and polystyrene (PS)/poly(methyl methacrylate) (PMMA)/poly(styrene-r-methyl methacrylate) (PS-r-PMMA), the polymer- of polystyrene (PS)/poly(methyl methacrylate) (PMMA)/tetrahydrofuran (THF), and the polymer-additive mixture polystyrene (PS)/poly(phenylene oxide) (PPO)/dimethyl phthalate (DMP) is demonstrated through comparison with reported experimental phase diagrams. An expression is derived for the effective interaction energy parameter, øeff, relevant to øAB values obtained from scattering measurements; values for øeff are computed and compared with results from experiments. The model provides a first iteration toward a simple and practical predictive tool that can solve common problems in polymer compounding, for example, choosing an effective compatibilizer for an immiscible polymer blend.

Introduction easier but limited in their generality. Tompa3 analyzed the phase relationships for systems of two polymers and Most commercial products made of polymers are, in a solvent. Based on this approach, phase diagrams of fact, a mixture of more than one component and, in immiscible polymer pairs and a solvent were calculated many instances, of more than one polymer. Mixing using experimental values for the ø parameter obtained polymers and additives is essential to enable processing ij by different techniques, such as glass transition and and to obtain good product properties, yet the mixing cloud point measurements, with good agreement.4-11 behavior of multicomponent polymer blends is not well In the Flory-Huggins formalism, the binary interac- understood, and what little is known has been learned FH empirically. Moreover, little experimental data on mul- tion parameter, øij , is inversely proportional to tem- ticomponent polymer systems have been reported in the perature and independent of composition and literature compared with binary systems. Because of the and as such has been found insufficient to describe the full range of phase behavior observed for polymer blends time and expense involved in experimental determina- - tion of multicomponent phase diagrams, a model ca- and .12 18 Empirical expressions that render pable of predicting the miscibility of multicomponent øij a function of composition and with several adjustable parameters have been developed with polymer mixtures would provide a powerful tool for - academic and industrial researchers. partial success.4 11,16,19 The complexity of phase diagram Three-component systems have been modeled as a calculations using these types of expressions has led to first approach to calculating the phase behavior of more simplified attempts to predict the miscibility of multicomponent systems. Several works have been ternary systems wherein only the of mixing 20 based on the classical Flory-Huggins model for the free was considered, resulting in limited predictive capabil- energy of mixing,1 which for three components can be ity. Even employing a modified interaction parameter - expressed as derived from experiments, the Flory Huggins model does not always yield an accurate description of the system phase behavior.21 ∆g φ φ φ mix ) A + B + C + In its original formulation, the Flory-Huggins model ln φA ln φB ln φC kT NAνA NBνB NCνC is a regular solution model that neglects the compress- ibility of the mixture components. Although it captures φAφB FH φAφC FH φBφC FH ø + ø + ø (1) essential features of upper critical solution transition ν AB ν AC ν BC AB AC BC (USCT) behavior, where component miscibility de- creases as temperature decreases, it cannot describe where ∆gmix is the free energy of mixing per unit lower critical solution transition (LCST) behavior, i.e., volume, φi is the of component i, Ni is with increasing temperature, which the number of segments per molecule, νi is the segmen- is found commonly in polymer-solvent systems and in FH - 14,17,18 tal volume, and νij and øij are the average segmental select polymer polymer systems. Other models volume and interaction parameter between species i and that account for the compressibility of multicomponent j, respectively. Scott2 calculated the phase equilibria in polymer systems have been successful in predicting a three-component system between a polymer, a solvent, LCST behavior, such as the Sanchez-Lacombe lattice and a nonsolvent for hypothetical components using fluid (LF) equation of state.22,23 In the LF model, different approximations to the equilibrium criteria such compressibility is accounted for by the addition of vacant as the “single approximation” and “complete sites into the lattice, which are assumed to mix ran- immiscibility approximation”, which make calculations domly with the polymer. The LF model has been applied

10.1021/ma0209803 CCC: $25.00 © 2003 American Chemical Society Published on Web 03/13/2003 Macromolecules, Vol. 36, No. 7, 2003 Ternary Polymer Mixtures 2509

21,24 to ternary systems and shown good agreement with xν ν experimental measurements. However, these calcula- FH ) A B - 2 ø (δA,0 δB,0) (3) tions generally employ an experimentally derived in- kT teraction parameter, which limits the usefulness of this kind of calculation as a practical predictive tool for Note that this term is always positive, destabilizing the systems that have not yet been explored experimentally. mixture. The third term of the model, which arises Several other equations of state that describe the solely from the compressibility of the components, can P-V-T behavior of polymers have been developed,25 be either positive or negative, enabling the model to such as the Flory-Orwoll-Vrij model and the Prigogine predict USCT and/or LCST behavior. Using eq 2, “square-well” cell model, but have not been applied to Ruzette and co-workers were able to qualitatively predict the phase behavior of over 30 binary polymer multicomponent polymer mixtures. Other models based 33,34 on perturbation theory, such as the perturbed-chain mixtures. 26 The compressible regular solution (CRS) model can statistical associating fluid theory (PC-SAFT) and the - perturbed hard-sphere chain equation (PHSC),27 have be compared to the Sanchez Lacombe LF model, which been successful in capturing the phase behavior en- provides a formal equation of state for polymer/polymer countered in polymer solutions. These methods treat and polymer/solvent mixtures. Both models assume a polymers as chains of freely joined spherical segments, mean field and account for the interaction energy by with interactions between different species divided into counting pairwise interactions. In contrast to the LF repulsive and attractive parts. These approaches still model, which accounts for thermal expansion by the require parameter fitting from experimental data; how- inclusion of holes as an extra component, the CRS model ever, a procedure based on the Flory-Huggins model calculates as the ratio of free to hard-core 28 volumes. The LF expression for an adimensional free and the UNIQUAC group contribution method has 35 been used to predict the enthalpy of mixing for a energy for a single component is multicomponent polymer mixture based on the chemical G structure of the components. Other models that utilize ) G˜) group contribution methods, such as the entropic free rNc* 1 F˜ volume activity coefficient model and the Holten- -F˜ + P˜ ν˜ + T˜ (ν˜ - 1) ln(1 -F˜) + ln (4) Anderson et al. equation of state, have yielded predic- [ r (ω)] tions of the phase behavior of polymers mixtures with partial success.29 where r is the number of repeat units in the polymer Atomistic approaches have also been used to model chain, Nc is the number of chains, * is the total the P-V-T behavior of polymers. In particular, the interaction energy per repeat unit, and ω is the number polymer reference interaction site model (PRISM) in- of configurations available to a r-mer in the close-packed tegral theory30,31 has been successful in simulating pure state, which is obtained from symmetry and polymer properties such as isothermal compressibilities flexibility parameters that are characteristic for the 36 ˜ ˜ F and cohesive energy densities. Nevertheless, its applica- component. T, P, ˜, and ν˜ are reduced temperature, tion to multicomponent polymer mixtures remains a pressure, , and volume defined as challenge. Utilizing novel Monte Carlo techniques such ˜)Tk ˜)Pν* ) 1 ) V as the Gibbs ensemble, Nath et al. simulated the T , P , ν˜ F (5) solubility of a mixture in a polymer,32 providing a * * ˜ Ncrν* first step toward the simulation of phase behavior of multicomponent mixtures including polymeric compo- where ν* is a close-packed volume. In a similar fashion nents. the CRS model for the free energy for a single compo- nent can be expressed in terms of reduced parameters A new model for polymer mixtures that takes into as follows: account compressibility and requires only pure compo- nent parameters as input was recently developed by ˜ F Gi Ti ˜ i Ruzette et al.33,34 For binary compressible polymer ) G˜ )-F˜ + P˜ ν˜ + ln (6) 2 i i i i N (F˜ - 1) blends the free energy of mixing per unit volume is niNiνiδi,0 i i expressed as with T˜ i, P˜ i, F˜i, and ν˜i defined as F F φA ˜ A φB ˜ B ∆g ) kT ln φ + ln φ + Tk P 1 V mix [N A N B] T˜ ) , P˜ ) , ν˜ ) ) (7) AνA BνB i 2 i 2 i F ν δ δ ˜ i niNiνi F F - 2 + F -F 2 - 2 i i,0 i,0 φAφB ˜ A ˜ B(δA,0 δB,0) φAφB(˜ A ˜ B)(δA δB ) (2)

where ni is the number of molecules present. The where F˜i is the reduced density (density/hard core entropy contribution, in both cases, has an inverse density), νi is the hard-core molar volume, δi,0 and δi dependence with molecular weight (r in LF, Ni in CRS) are the solubility parameters at 0 K and temperature but for the CRS model approaches zero as the size of T, respectively, Ni is the number of repeat units per the molecule increases, versus a finite value for the LF chain, and φi is the volume fraction of component i. The model. Equation 6 provides an equation of state analo- standard Berthelot mixing rule is invoked in this model, gous to the van der Waals’ equation of state. such that the solubility parameter of the mixed state is For mixtures, the LF model in its widely used original a geometric average of the component values. The first formulation35 assumes that every component in the term of eq 2 accounts for the translational entropy of mixture, including the holes, has the same lattice mixing in a similar way to the Flory-Huggins formal- volume (per segment). Solving the phase equilibrium ism. The second term can be related directly to the in mixtures thus requires rescaling of variables to Flory-Huggins interaction parameter approximation: account for components of disparate sizes,36 which 2510 Gonzalez-Leon and Mayes Macromolecules, Vol. 36, No. 7, 2003

the density at a given temperature and pressure, Fi,to F/ the hard-core density, i , and quantifies the space occupied by the hard-core volume of the molecule: F V F˜ ) i ) hc,i (10) i F/ V i i

In a similar fashion, a relation for the total reduced density, F˜, can be written as

F V F˜ ) ) hc (11) F* V Figure 1. Chemical structures for PC, TMPC, PPO, and DMP. Equation 8 can be expressed using reduced densities, F˜ , and volume fractions, φ ,as complicates the analysis. In the CRS model, the spinodal i i compositional boundary is described by a simple quad- p p 33 ∆Scomb 1 -F˜ ratic expression with an analytical solution. The )-∑n ln(φ ) + ∑n ln (12) advantage of the latter over more elaborate EOS i i i ( ) k i i 1 -F˜ models22,36 appears to be its simplicity for predicting i phase behavior of polymer mixtures. The second term in this expression is typically orders In the present work, we extend the binary compress- of magnitude smaller than the first term and can ible regular solution model to multicomponent systems generally be neglected.33 and calculate spinodal curves for several ternary poly- The total change of the interaction energy upon mer mixtures, including the ternary polymer blends of mixing can be simply derived as tetramethylpolycarbonate (TMPC), polycarbonate (PC), shown in Figure 1, with poly(styrene-r-acrylonitrile) ∆E ) E - E (13) (SAN) and polystyrene (PS), poly(methyl methacrylate) mix mixed pure (PMMA) with poly(styrene-r-methyl methacrylate) (PS- The energy of the pure state, E , is calculated by r-PMMA) for comparison with experimentally reported pure 24,37,38 counting the number of pairwise interactions of each behavior. Calculations for a mixture of two in- component in its pure state, assuming a mean-field compatible polymers with a common solvent are also approximation. Defining the hard-core cohesive energy presented, namely, polystyrene (PS)/poly(methyl meth- density of compound i as acrylate) (PMMA)/tetrahydrofuran (THF).5 Finally, the spinodal diagram for a mixture of a compatible polymer z 2 )-1 ii blend, polystyrene (PS)/poly(phenylene oxide) (PPO), δi,0 (14) and a plasticizer, dimethyl phthalate (DMP), is pre- 2 νi sented. The objective of the present work is to demon- strate the use of the model for phase behavior predic- where ii is the segmental interaction energy of compo- tions of ternary systems, highlighting its successes and nent i with itself and z is the number of nearest- shortcomings. neighbor segments (or solvent molecules) in the pure state, Epure becomes Multicomponent Model p p 1 zii niNiνi For the mixing of p different species with ni molecules ) )- 2F Epure ∑niNiνi( ) ∑niNiνiδi,0 ˜ i (15) and Ni segments of hard-core volume νi (volume at 0 K i 2 ν V i and zero pressure), the combinatorial entropy of mixing, i i 39 1 as presented by Hildebrand and Flory, can be ex- Here, for each component i, the energy of the hard-core pressed as state is diluted by a factor F˜i. p The energy in the mixed state can be calculated in a ∆Scomb Vf,m ) similar manner. In the mixed state the pairwise inter- ∑ni ln( ) (8) actions of each component with itself should be counted, k i V f,i plus the interactions between the different species. The number of each kind of interaction is proportional to where Vf,m is the total free volume present in the mixture and V is the free volume of component i in its the volume fraction of the components involved. As- f,i suming a geometric average for the cross-interaction pure state. These free volumes can be written as the 33 differences between the total volumes and the hard-core energy density volumes: 1 zxiijj ) - δ 2 ) ) δ δ (16) Vf,i Vi Vhc,i ij,0 2 i,0 j,0 xνiνj p ) - Vf,m V ∑Vhc,i (9) the total energy of the mixed state for the p components i can be expressed as where the hard-core volume of component i, Vhc,i,is p p ) ) - F simply expressed by Vhc,i niNiνi. These definitions can Emixed ∑∑ φjniNiνiδi,0δj,0 ˜ j (17) be related to a reduced density, F˜i, which is the ratio of i j Macromolecules, Vol. 36, No. 7, 2003 Ternary Polymer Mixtures 2511

2 Using eqs 15 and 17 in eq 13, we obtain the change in ∂ ∆g ∂2∆g interaction energy upon mixing: mix mix 2 2 2 ∂φi ∂φi ∂φj ∂ ∆gmix ∂ ∆gmix p p p ) - 2 2 2 2 2 ∂ g ) - F + F ∂ ∆gmix ∆ mix ∂φi ∂φj ∆Emix ∑∑ φjniNiνiδi,0δj,0 ˜ j ∑niNiνiδi,0 ˜ i (18) [ ] i j i 2 ∂φj ∂φi ∂φj 2 2 With eqs 12 and 18, an expression for the change in the ∂ ∆g ∂ ∆g mix mix ) free energy of mixing for compressible multicomponent 0 (24) ∂φi ∂φj ∂φj ∂φi systems per unit volume, ∆gmix, can be built:

p F p p where the i and j components could be any two of A, B, φi ˜ i ) + - F F + and C. ∆gmix kT∑ ln(φi) ∑∑ φj ˜ jφi ˜ iδi,0δj,0 N ν i i i i j Component Properties p F 2 2 The pure component properties needed for the calcu- ∑φi ˜ i δi,0 (19) i lation of the spinodal curves were obtained from previ- ously reported experimental P-V-T measurements23,25 For the case of three components, eq 19 becomes, after and by group contribution methods40,41 following the some algebraic manipulation, approach of Ruzette et al.33,34 The density dependence on temperature, Fi(T), was φ F˜ φ F˜ assumed to follow the form ) A A + B B + ∆gmix kT( ln(φA) ln(φB) NAνA NBνB / -R F )F iT F i(T) i e (25) φC ˜ C ln(φ ) + φ φ (F˜ δ -F˜ δ )2 + φ φ (F˜ δ - N ν C ) A B A A,0 B B,0 A C A A,0 C C where Ri is the coefficient of thermal expansion for F 2 + F -F 2 component i, obtained from exponential fits to previ- ˜ CδC,0) φBφC(˜ BδB,0 ˜ CδC,0) (20) ously reported P-V-T data,25 either as the empirical As shown in the binary case,33 this expression can be Tait equation or from the Sanchez-Lacombe lattice 23 F/ separated into compressible and incompressible contri- fluid equation of state. The hard core density, i ,is butions for each binary interaction: taken to be the extrapolation of this fit to 0 K. From / this value, one can obtain F˜ and ν ) M /N F , where ) i i i Av i ∆gmix Mi is the molecular weight of the repeat unit or molecule F F F φA ˜ A φB ˜ B φC ˜ C and NAv is Avogadro’s number. kT ln(φ ) + ln(φ ) + ln(φ ) + To obtain solubility parameter values for a given (N ν A N ν B N ν C ) A A B B C C temperature T, the solubility parameter at 298 K was F F - 2 + F -F 2 - 2 + 41 φAφB ˜ A ˜ B(δA,0 δB,0) φAφB(˜ A ˜ B)(δA δB ) first calculated by group contribution methods and 2 2 2 then extrapolated to the required temperature using the φ φ F˜ F˜ (δ - δ ) + φ φ (F˜ -F˜ )(δ - δ ) + A C A C A,0 C,0 A C A C A C following expression:33,34 φ φ F˜ F˜ (δ - δ )2 + φ φ (F˜ -F˜ )(δ 2 - δ 2) B C B C B,0 C,0 B C B C B C F (21) ˜ i(T) δ 2(T) ) δ 2(298) (26) i i (F˜ (298)) For the binary phase diagrams presented in this i paper, the binodal and spinodal curves were calculated For random copolymers, properties were estimated by using eq 21 and the classic equilibrium criteria for averaging the values of their homopolymer components. binodal and spinodal conditions. For the binodal, the Parameters used in the LF equation of state were chemical potentials of each component should be equal / / for every phase present: obtained using molar averages for PLF and TLF, the LF characteristic pressure and temperature, respectively, / 1 2 1 2 F µ - µ ) 0 and µ - µ ) 0 (22) while LF was calculated using a weight-based aver- A A B B age.42 k The values of the component parameters used for where µ is the of component i in i calculation of the binodal and spinodal diagrams pre- the phase k. The spinodal condition is reached by setting sented in this paper are listed in Table 1. the second derivative of the free energy of mixing with 33 respect to composition equal to zero: Phase Diagrams ∂2∆g Ternary Polymer Blends. The spinodal diagram for mix ) 0 (23) a ternary blend of tetramethylpolycarbonate (TMPC), 2 ∂φi polycarbonate (PC), and a random copolymer of styrene with acrylonitrile (14.7 wt % AN, SAN 14.7) was For the ternary systems studied here, only the spin- calculated from eq 24, at T) 413 K, as shown in Figure odal curves of each system are calculated. The spinodal 2. The molecular weights used in the calculation matched curves define the boundary between the stable and those in ref 24 (Mn,TMPC ) 27 800, Mn,PC ) 37 000, and metastable regions and should be sufficient to qualita- Mn,SAN14.7 ) 83 000 g/mol) to compare calculations with tively predict the phase behavior of the system. Follow- experimental cloud points. As can be seen, agreement ing Scott,2 for ternary systems the boundary points of with the previously reported experimental data is quite the spinodal curve are found by the solution to the good, despite the simplifying assumptions of the model. following equation: Binary phase diagrams for this ternary system were 2512 Gonzalez-Leon and Mayes Macromolecules, Vol. 36, No. 7, 2003

Table 1. Parameters Used in Calculations

F* R δ(298) NAvν (g/cm3) (10-4 K-1) (J1/3/cm3/2) (cm3/mol) PS 1.24 5.13 18.19 83.87 PMMA 1.43 5.61 19.65 69.83 PS-r-PMMAa 1.33 5.58 18.85 76.20 PC 1.51 6.33 19.07 167.73 TMPC 1.44 7.00 19.68 215.12 SAN 14.7 1.29 5.73 19.44 70.22 THFb 1.36 14.6 18.6 52.82 PPO 1.45 7.23 18.90 83.04 DMPc 1.49 7.66 20.44 129.82 a R calculated from the characteristic properties of PS and PMMA reported in ref 23; S:MMA molar ratio 50:50. b Solubility parameter from Barton.64 c P-V-T properties estimated using group contribution methods from Reid et al.65

Figure 4. Calculated TMPC/PC phase diagram for MTMPC ) 27 800 and MPC ) 37 000 g/mol. Spinodal (- - -) and binodal (s) boundaries are shown.

Figure 2. Spinodal diagram at T ) 413 K of TMPC/PC/SAN 14.7 for MTMPC ) 27 800, MPC ) 37 000, and MSAN 14.7 ) 83 000 g/mol. Open squares (0) are miscible compositions; filled circles (b) are compositions where two phases exist. Experimental data from Kim et al.24 for a system with equivalent reported Mn.

Figure 5. Calculated phase diagram for PC/SAN 14.7 for MPC ) 37 000 and MSAN 14.7 ) 83 000 g/mol. Spinodal (- - -) and binodal (s) boundaries are shown.

observed LSCT behavior. Assuming a polydispersity (Mw/Mn) of the TMPC equal to 2, implying Mn,TMPC ) 16 500 g/mol, the same calculation using MTMPC ) 16 500 g/mol gives close agreement with experimental values. The second binary system, TMPC/PC, is shown in Figure 4 with MTMPC ) 27 800 and MPC ) 37 000 g/mol. Although no phase diagram is reported in the literature for this system, it has been noted that these polymers remain miscible up to their degradation tem- perature,24 consistent with the USCT behavior predicted for this mixture. Finally, in Figure 5 the binary phase ) Figure 3. Calculated phase diagram of TMPC and SAN 14.7 diagram for PC and SAN 14.7 is shown for MPC 37 000 g/mol and M ) 83 000 g/mol. It has been previ- for MSAN 14.7 ) 83 000 and MTMPC ) 16 500 g/mol (a) and MTMPC SAN14.7 ) 33 000 g/mol (b). Spinodal (- - -) and binodal (s) boundaries ously reported44 that mixtures of PC and SAN are highly are shown. Experimental cloud points (9) for Mw,TMPC ) 33 000 immiscible, and this trend is again borne out by the ) 43 and Mn,SAN 14.7 83 000 g/mol taken from Kim and Paul. calculations. The ternary spinodal diagram at 298 K was also also calculated. The phase diagram for TMPC/SAN 14.7 calculated for a mixture of PS, PMMA, and a random is shown in Figure 3, using MTMPC ) 33 000 and MSAN14.7 copolymer of PS and PMMA with a 50:50 S:MMA molar ) 83 000 g/mol to compare with experimental cloud ratio. The molecular weights used for the calculation of 43 points. For a system in which Mw,TMPC ) 33 000 g/mol, Figure 5 were MPS ) 10 400, MPMMA ) 10 000, and the computed phase diagram qualitatively captures the MPS-r-PMMA ) 10 200 g/mol. Figure 6 indicates that the Macromolecules, Vol. 36, No. 7, 2003 Ternary Polymer Mixtures 2513

Figure 6. PS/PMMA/PS-r-PMMA at 298 K diagram for M Figure 7. PS/PMMA/THF spinodal diagram at T ) 298 K PS ) ) ) 10 400, MPMMA ) 10 000, and MPS-r-PMMA ) 10 200 g/mol for MPS 13 000 and MPMMA 32 800 g/mol. Experimental 0 5 showing spinodal boundary. cloud points ( ) are shown for a system of equivalent Mn. copolymer, as might be expected, improves the miscibil- ity between PS and PMMA. It has been reported37,38 that for this specific system PMMA is more miscible with the copolymer than is PS. In our calculation, however, the opposite is observed. As can be seen in the diagram, the immiscibility region is closer to the PMMA/ PS-r-PMMA boundary, indicating a greater incompat- ibility between this pair. The discrepancy between experiment and calculation points to an apparent failure of the model’s geometric averaging of the A-B interac- tion energy. In this system PMMA is the high-density component and also has a higher cohesive energy density than PS (Table 1). Upon mixing, PMMA self- interactions are diluted as a consequence of compress- ibility, since the density of the mixture is equal to an average of the component densities. On the other hand, PS will increase its density by mixing, increasing the magnitude of self-interactions. This densification pro- Figure 8. PS/PPO/DMP spinodal diagram at T ) 423 K for ) ) cess favors the mixing of PS/PS-r-PMMA compared to MPS 104 000 and MPPO 120 000 g/mol. PMMA/PS-r-PMMA, giving rise to the asymmetry ob- served in the calculated spinodal diagram in Figure 6. Ternary Mixture of Two Homopolymers and a Experimentally, this effect must be offset by a styrene- Plasticizer. Finally, an example of the possible ap- methyl methacrylate interaction that is not simply the plication of the model to polymer processing and com- geometric average of the component self-interactions. pounding problems is presented. The spinodal diagram The inability to capture this feature of the ternary for a system of two homopolymers and a plasticizer was mixture points to the need for an alternate means to calculated. Plasticizers are typically or low compute A-B interactions for the model. that are added to polymers to improve Ternary Mixture of Two Homopolymers and a their processability and modify their mechanical prop- Solvent. As a third illustration of the ternary mixture erties. The spinodal diagrams at 298 and 423 K for a model, we calculated spinodal diagrams for two incom- mixture of polystyrene (PS)/poly(phenylene oxide) (PPO) patible polymers, PS and PMMA, and a common sol- and the plasticizer dimethyl phthalate (DMP) with MPS vent, tetrahydrofuran (THF). Figure 7 shows a spinodal ) 104 000 and MPPO ) 120 000 g/mol were calculated. diagram for T ) 298 K for PS, PMMA, and THF, where The PS/PPO blend is reported in the literature to be the molecular weights of the polymers used for the totally miscible.45 In our calculations, the ternary calculations matched those in a previously studied system is totally compatible at room temperature, 5 system (Mn,PS ) 13 000, Mn,PMMA ) 32 800 g/mol). Here showing that DMP is a compatible plasticizer for the no modification to the model was made due to the polymer blend. However, Figure 8 predicts that, at the presence of the solvent; the solvent molecular weight higher required for melt processing, the used was for one THF molecule (72 g/mol), while NTHF system becomes partially immiscible and a two-phase ) 1. As can be seen in Figure 7, the general thermody- zone appears. namic behavior of the system is well captured. As in The spinodal diagrams calculated in this work suggest other systems comprising two incompatible polymers that the generalized model for compressible regular and a solvent,4,6 miscibility is not achieved unless the solutions is a promising direction toward the prediction amount of solvent present is significant (more than of multicomponent polymer mixture phase behavior. It ∼70%), where translational entropy gains due to the should be noted, however, that not all ternary systems presence of the small molecules dominate the free are well predicted by our model. In the PS/TMPC/PC energy. system, the model failed to predict the reported phase 2514 Gonzalez-Leon and Mayes Macromolecules, Vol. 36, No. 7, 2003

Table 2. Interaction Parameter Comparison

A(CRS) B(CRS) øeff(CRS) A(exp) B(exp) øeff(exp) PMMA/PS -0.067 28.9 0.001 0.028 3.90 0.037 PS/PS-r-PMMA -0.018 12.4 0.008 0.012 PS/PPO -0.033 -3.76 -0.042 0.121 -78.0 -0.063 TMPC/PS 0.199 -50.5 0.080 0.110 -54.3 -0.018 Analogously, the second derivative of eq 2 with respect to provides an expression for 1/I(0) for a compressible binary blend: F˜ F˜ 2F˜ F˜ 1 ) -2 A + B - A B - 2 - kn [ (δA,0 δB,0) I(0) NAφAνA NBφBνB kT 2 (F˜ -F˜ )(δ - δ ) (29) kT A B A,0 B,0 ] Substituting the approximation F ≈ -R ˜ i 1 iT (30)

an equivalence for øeff can be extracted from eq 29, if we assume the form for 1/I(0) given in eq 27: R R ) - A - B - 2 R 2 + øeff ν0[( ) ( AδA,0 φANAνA φBNBνB k Figure 9. Calculated PS/TMPC phase diagram for M ) R 2 -R -R + 1 2 + W,PS BδB,0 AδA,0δB,0 BδA,0δB,0) (δA,0 318 000 and MW,TMPC ) 52 600 g/mol. Spinodal (- - -) and kT binodal (s) boundaries are shown as well as experimental 2 T 2 2 2 2 9 46 δ - 2δ δ ) + (R δ +R δ - LCST cloud points ( ). B,0 A,0 B,0 k A A,0 B B,0 diagram for the PS/TMPC pair46 where no UCST R R 2 A BδA,0δB,0)] (31) behavior is present in the studied temperature window. Nevertheless, the LCST behavior for the system is From eq 31, one may note that the effect of fitting captured at higher temperatures as shown in Figure 9. scattering data from a compressible binary blend to the For this system, the solubility parameter of TMPC had form for 1/I(0) in eq 27 naturally gives rise to an to be approximated because precise group contributions “entropic component” of øeff that is concentration and needed for this component were not available. In molecular weight dependent. The temperature depen- practice, we have found the model to be quite sensitive dence of øeff is additionally found to be more complex to δ values, so that inaccuracies in the calculation of δ than the classical 1/T behavior.13,16,57 might account for the disagreement. To compare the CRS model results with experimen- We found also discrepancies in systems that are tally obtained øAB expressions of the form shown in eq reported to be totally miscible,47-49 such as, poly- 28, we calculated øeff from eq 31 and subsequently fit (hydroxy ether) of bisphenol A/PMMA/poly(ethylene this “data” plotted as øeff vs 1/T to a straight line, oxide) and poly(hydroxy ether) of bisphenol A/poly(vinyl yielding values for A and B from the y-intercept and methyl ether) (PVME)/poly(epichlorohydrin), for which slope, respectively. For systems for which øAB is reported strong specific interactions exist between polymer pairs. for a single temperature only, eq 31 was used to com- In our model specific interactions such as hydrogen pute øeff directly. Results for binary mixtures related bonding are not properly captured by eq 16, and such to this work for which experimental data were avail- systems are thus predicted to be incompatible. able are given in Table 2, along with values obtained A further evaluation of the model can be made by from scattering experiments.50-54 (Note that deutera- comparing model predictions for the interaction energy tion was not accounted for in the calculations.) The øeff with experimentally determined values for ø reported AB for PS/PMMA was calculated using MPS ) 5470 and from scattering data on binary blends.50-54 In analyzing MPMMA ) 5260 g/mol, at the composition PS ) 44 wt % such data, the Flory-Huggins model is often assumed, and 150 °C. The øeff so obtained is compared with values whereby the inverse scattering intensity extrapolated for an equivalent block copolymer from the literature to zero wavevector, 1/I(0), can be fit to the expression55,56 51 (Mn,PS-b-PMMA ∼ 27 000 g/mol), capturing the reported positive value for the interaction parameter but with a 2ø 1 ) -2 1 + 1 - AB difference in magnitude and temperature dependence. kn [ ] (27) ) For PS/PS-r-PMMA, øeff was calculated using MPMMA I(0) NAφAνA NBφBνB νAB MPS-r-PMMA) 87 000 g/mol, a 50:50 random copolymer of styrene and MMA, and a mixture composition of and øAB often obeys the form φPMMA ) 0.2. The øeff was calculated at 200 °C and is close to the reported value from the literature.52 For the B ) + PS/PPO system, øeff was calculated using MPS ) 32 380 øAB A (28) T and MPPO ) 16 816 g/mol with a PS weight percent of 98%. For this extreme composition, the CRS model 2 56 In eq 27, kn is the scattering contrast factor. captures correctly the negative sign of øeff and also Macromolecules, Vol. 36, No. 7, 2003 Ternary Polymer Mixtures 2515 predicts a high sensitivity to molecular weight and (21) Brannock, G. R.; Paul, D. R. Macromolecules 1990, 23, 5240. (22) Sanchez, I. C.; Lacombe, R. H. Macromolecules 1978, 11, composition for the øeff temperature dependence, chang- > 1145. ing sign of the slope for compositions 3.5 wt % PS (23) Sanchez, I. C.; Panayiotou, C. G. In Models for Thermody- 53 (experimental composition range ∼ 1-3%). For the namic and Phase Equilibria Calculations; Sandler, S. I., Ed.; TMPC/PS system, MPS ) 25 000 and MTMPC ) 46 000 Marcel Dekker: New York, 1993. g/mol, and a PS weight percent of 25% was used for the (24) Kim, C. K.; Kim, J. J.; Paul, D. R. J. Polym. Eng. Sci. 1994, 34, 1788. calculation of øeff at 150 °C. Although the temperature (25) Rodgers, P. A. J. Appl. Polym. Sci. 1993, 48, 1061. dependence is quite well captured by the CRS model, (26) Gross, J.; Sadowski, G. Ind. Eng. Chem. Res. 2002, 41, 1084. 54 the reported negative sign of øeff is not obtained, (27) Cardelli, C.; Conto, G.; Gianni, P.; Porta, R. J. Therm. Anal. consistent with the discrepancies found in the phase Calorim. 2000, 62, 135. diagram for this system shown in Figure 9. We will (28) Pappa, G. D.; Kontogeorgis, G. M.; Tassios, D. P. Ind. Eng. Chem. Res. 1997, 36, 5461. explore such comparisons in more depth in a future (29) Favari, F.; Bertucco, A.; Elvassore, N.; Fermeglia, M. Chem. article. Eng. Sci. 2000, 55, 2379. (30) Schweizer, K. S.; David, E. F.; Singh, C.; Curro, J. G.; Conclusions Rajasekaran, J. J. Macromolecules 1995, 28, 1528. (31) Chatterjee, A. P.; Schweizer, K. S. Macromolecules 1998, 31, The free energy model presented herein shows quali- 2353. tative capability in predicting the phase behavior of (32) Nath, S. K.; Banaszak, B. J.; de Pablo, J. J. Macromolecules multicomponent polymer systems using only pure com- 2001, 34, 7841. (33) Ruzette, A.-V. G.; Mayes, A. M. Macromolecules 2001, 34, ponent properties. The model shows to be useful not only 1894. for multicomponent polymer mixtures but for polymer (34) Ruzette, A.-V. G.; Banerjee, P.; Mayes, A. M.; Russell, T. P. /solvent and polymer /additive mixtures as well, captur- J. Chem. Phys. 2001, 114, 8205. ing their general thermodynamic behavior. Further (35) Sanchez, I. C.; Lacombe, R. H. J. Phys. Chem. 1976, 80, 2352. (36) Lacombe, R. H.; Sanchez, I. C. J. Phys. Chem. 1976, 80, 2568. modifications to the model are clearly needed, as il- (37) Winey, K. I.; Berba, M. L.; Galvin, M. E. Macromolecules lustrated by its failure to capture the asymmetry 1996, 29, 2868. experimentally observed in the PS/PMMA/PS-r-PMMA (38) Kulasekere, R.; Kaiser, H.; Anker, J. F.; Russell, T. P.; Brown, system and the low-temperature immiscibility of PS/ H. R.; Hawker, C. J.; Mayes, A. M. Macromolecules 1996, 29, 5493. TMPC. Extension of the model to account for specific (39) Hildebrand, J. H. J. Chem. Phys. 1947, 15, 225. A-B interactions and to include block copolymer com- (40) Boudouris, D.; Constantinou, L.; Panayiotou, C. Ind. Eng. ponents is now underway and should greatly expand Chem. Res. 1997, 36, 3968. the application of this approach to industrially relevant (41) Van Krevelen, D. W.; Hoftyzer, P. J. Properties of Polymers. blending problems.47-49,58-63 Correlation with Chemical Structure; Elsevier: New York, 1972. (42) Panayiotou, C.; Oehmke, F. Fluid Phase Equilib. 1996, 126, Acknowledgment. This work was supported by 289. matching grants from the Lord Foundation and the Lord (43) Kim, C. K.; Paul, D. R. Polymer 1992, 33, 2089. Corporation and by the Seaver Institute. (44) Li, H.; Yang, Y.; Fujitsuka, R.; Ougizawa, T.; Inoue, T. Polymer 1999, 40, 927. (45) Lutz, J. T. Thermoplastic Polymer Additives, Theory and References and Notes Practice; Marcel Dekker: New York, 1989. (46) Guo, W.; Higgins, J. S. Polymer 1990, 31, 699. (1) Flory, P. J. Principles of Polymer Chemistry; Cornell Univer- (47) Robeson, L. M.; Hale, W. F.; Merriam, C. N. Macromolecules sity Press: Ithaca, NY, 1953. 1981, 14, 1644. (2) Scott, R. L. J. Chem. Phys. 1949, 17, 268. (48) Hu, Y. H.; Painter, P. C.; Coleman, M. M. J. Appl. Polym. (3) Tompa, H. Trans. Faraday Soc. 1949, 45, 1142. Sci. 1998, 70, 1265. (4) Van Den Esker, M. W.; Vrij, A. J. Polym. Sci. 1976, 14, 1943. (49) Min, K. E.; Chiou, J. S.; Barlow, J. W.; Paul, D. R. Polymer (5) Geveke, D. J.; Danner, R. P. J. Appl. Polym. Sci. 1993, 47, 1987, 28, 1721. 565. (50) Balsara, N. P. In Physical Properties of Polymers Handbook; (6) Tong, Z.; Einaga, Y.; Kitagawa, T.; Fujita, H. Macromolecules Mark, J. E., Ed.; AIP Press: New York, 1996. 1989, 22, 450. (51) Russell, T. P. Macromolecules 1993, 26, 5819. (7) Lee, S. C.; Woo, E. M. J. Polym. Sci., Part B 2002, 40, 747. (52) Hellmann, E. H.; Hellmann, G. P. Macromol. Chem. Phys. (8) Pomposo, J. A.; Mugica, A.; Areizaga, J.; Cortazar, M. Acta 1997, 198, 329. Polym. 1998, 49, 301. (53) Maconnachie, A.; Kambour, R. P.; White, D. M.; Rostami, S.; (9) Clark, A. H. Carbohydr. Polym. 2000, 42, 337. Walsh, D. J. Macromolecules 1984, 17, 2645. (10) Favre, E.; Mguyen, Q. T.; Clement, R.; Neel, J. Eur. Polym. (54) Kim, E.; Kramer, E. J.; Osby, J. O.; Walsh, D. J. J. Polym. J. 1996, 32, 303. Sci., Part: Polym. Phys. 1995, 33, 467. (11) Anderko, A. Int. J. Thermophys. 1994, 15, 1221. (55) Frielinghaus, H.; Schwahn, D.; Willner, L. Macromolecules (12) Balsara, N. P.; Fetters, L. J.; Hadjichristidis, N.; Lohse, D. 2001, 34, 1751. J.; Han, C. C.; Graessley, W. W.; Krishnamoorti, R. Macro- (56) Higgins, J. S.; Benoit, H. C. Polymers and Neutron Scattering; molecules 1992, 25, 6137. Oxford University Press: New York, 1994. (13) Krishnamoorti, R.; Graessley, W. W.; Balsara, N. P.; Lohse, (57) Koningsveld, R.; Stockmayer, W. H.; Nies, E. Polymer Phase D. J. J. Chem. Phys. 1994, 100, 3894. Diagrams; Oxford University Press: New York, 2001. (14) Krishnamoorti, R.; Graessley, W. W.; Fetters, L. J.; Garner, (58) Qipeng, G. Eur. Polym. J. 1990, 26, 1329. R. T.; Lohse, D. J. Macromolecules 1995, 28, 1252. (59) Espi, E.; Fernandez-Berridi, M. J.; Iruin, J. J. Polym. Eng. (15) Frielinghaus, H.; Schwahn, D.; Willner, L. Macromolecules Sci. 1994, 34, 1314. 2001, 34, 1751. (60) Zhang, H.; Bhagwagar, D. E.; Graf, J. F.; Painter, P. C.; (16) Klein, J.; Kerle, T.; Zink, F.; Eiser, E. Macromolecules 2000, Coleman, M. M. Polymer 1994, 35, 25. 33, 1298. (61) Shull, K. R.; Kramer, E. J. Macromolecules 1990, 23, 4769. (17) Chen, Y. Y.; Lodge, T. P.; Bates, F. S. J. Polym. Sci. 2000, (62) Lyatskaya, Y.; Gersappe, D.; Balazs, A. C. Macromolecules 38, 2965. 1995, 28, 6278. (18) Hao, W.; Elbro, H. S.; Alessi, P. Polymer Solution Data (63) Noolandi, J.; Hong, K. M. Macromolecules 1982, 15, 482. Collection Part 2: Solvent Activity coefficients at Infinite (64) Barton, A. F. M. Handbook of Solubility Parameters and , Part 3: Liquid-Liquid Equilibrium; DECHEMA Other Cohesion Parameters, 2nd ed.; CRC Press: Boca Raton, Chemistry Data Series: Frankfurt Main, Germany, 1992. FL, 1991. (19) Caibao, Q.; Mumby, S. J.; Eichinger, B. E. Macromolecules (65) Reid, R. C.; Prausnitz, J. M.; Poling, B. E. The Properties of 1991, 24, 1655. & Liquids, 4th ed.; McGraw-Hill: New York, 1987. (20) Nishimoto, M.; Keskkula, H.; Paul, D. R. Polymer 1989, 30, 1279. MA0209803