Vous êtes sur la page 1sur 50

966

Ind. Eng. Chem. Res. 1997, 36, 966-1015

Mathematical Modeling of Multicomponent Chain-Growth Polymerizations in Batch, Semibatch, and Continuous Reactors: A Review
Marc A. Dube
Department of Chemical Engineering, University of Ottawa, Ottawa, Ontario, Canada K1N 6N5

Joao B. P. Soares and Alexander Penlidis*


Department of Chemical Engineering, University of Waterloo, Waterloo, Ontario, Canada N2L 3G1

Archie E. Hamielec
McMaster Institute for Polymer Production Technology, Department of Chemical Engineering, McMaster University, Hamilton, Ontario, Canada L8S 4L7

A practical methodology for the computer modeling of multicomponent chain-growth polymerizations, namely, free-radical and ionic systems, has been developed. This is an extension of a paper by Hamielec, MacGregor, and Penlidis (Multicomponent free-radical polymerization in batch, semi-batch and continuous reactors. Makromol. Chem., Macromol. Symp. 1987, 10/11, 521). The approach is general, providing a common model framework which is applicable to many multicomponent systems. Model calculations include conversion of the monomers, multivariable distributions of concentrations of monomers bound in the polymer chains and molecular weights, long- and short-chain branching frequencies, chain microstructure, and crosslinked gel content when applicable. Diffusion-controlled termination, propagation, and initiation reactions are accounted for using the free-volume theory. When necessary, chain-lengthdependent diffusion-controlled termination may be employed. Various comonomer systems are used to illustrate the development of practical semibatch and continuous reactor operational policies for the manufacture of copolymers with high quality and productivity. These comprehensive polymerization models may be used by scientists and engineers to reduce the time required to develop new polymer products and advanced production processes for their manufacture as well as to optimize existing processes.
1. Introduction Mathematical models and their role in science/ engineering are points of constant debate, especially when models are employed in an industrial environment. The role of a mathematical model is often misinterpreted; as a result, we frequently blame the model, instead of blaming our own lack of understanding about a process as well as our reluctance to experiment with a process in a meaningful and systematic way. Why, then, are models useful? 1. Models enhance our process understanding since they direct further experimentation. They act as the reservoir of ones knowledge about a process, and hence they may reveal interactions in a process that may be difficult, if not impossible, to visualize/predict solely from memory or experience, especially when many factors vary simultaneously. Since a model is a concise, compact form of process knowledge, models enhance transferability of knowledge; they may act eventually as an inference engine, closely resembling the train of thought of an experienced human. In a sense, mathematical modeling is the best way to find out what one does not know about a process! 2. Models are useful for process design, parameter estimation, sensitivity analysis, and process simulation.
* To whom correspondence should be addressed. Phone: (519) 888-4567. Fax: (519) 746-4979. E-mail: penlidis@ cape.uwaterloo.ca. E-mail: dube@genie.uottawa.ca. S0888-5885(96)00481-2 CCC: $14.00

The significance of these is quite obvious. A valid model allows one to test deviations from process trajectories using a simulator in lieu of running experiments. Cost effectiveness implications are also obvious. 3. Models are useful for process optimization, especially when dealing with highly nonlinear problems such as grade changes/switchovers in batch, semibatch, and continuous reactors. Extensions to recipe modifications and design are another application. 4. Models are useful for safety/venting considerations. It is very useful to be able to extrapolate to different operating conditions and anticipate worst-case scenarios or investigate the possible effects of process factors. In this case one may be better prepared to tackle situations that might not always be apparent from the outset. 5. Models are useful for optimal sensor selection and testing, sensor location, filtering and inference of unmeasured properties, and process control. The trends nowadays in process control are toward model-based control, and as the term signifies, application of advanced control techniques may not be possible without a model. 6. Finally, since a model contains process knowledge and is transferable, interactive models are extremely useful for the education and training of new (and old) personnel. In this paper, a practical methodology for the computer modeling of multicomponent chain-growth polymerizations, namely, free-radical and ionic systems, is developed. This is an extension of the paper by Hamielec
1997 American Chemical Society

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 967

et al. (1987). The approach is general, providing a common model framework which is applicable to many multicomponent systems. Various comonomer systems are used to illustrate the development of practical reactor operational policies for the manufacture of polymers with high quality and productivity. Thorough general reviews of polymerization reactor modeling have recently been published (Penlidis et al., 1985b; Hamielec et al., 1987; Rawlings and Ray, 1988). The model developed in this paper employs the pseudokinetic rate constant method (Hamielec et al., 1987; Tobita and Hamielec, 1991; Xie and Hamielec, 1993a,b). Our development uses many techniques and ideas similar to those contained in Hoffman (1981), Broadhead et al. (1985), Penlidis et al. (1986), Hamielec et al. (1987), Mead and Poehlein (1988, 1989a), Rawlings and Ray (1988), Maxwell et al. (1992b), Fontenot and Schork (1992-93a,b), Xie and Hamielec (1993a,b), Casey et al. (1994), and Urretabizkaia and Asua (1994). 2. Model Development: Free-Radical Polymerizations The objective of this section is to define the equations which form a mechanistic model to simulate bulk (or suspension), solution, and emulsion free-radical homopolymerization, copolymerization, and multicomponent polymerization (three or more monomer types) in wellstirred batch, semibatch, and continuous modes. The equations presented in this section are valid for the general case of an unsteady-state CSTR. For a CSTR operating at steady state, the accumulation derivative terms can be set equal to zero to give a set of algebraic equations. For a semibatch reactor, the outflow terms should be eliminated, and for a strictly batch reactor, all inflow and outflow terms should be eliminated. However, it is usually advantageous to consider complete equations since one has the flexibility of handling all these reactor situations with a single model. The model is comprised of a set of mathematical expressions which describe the physical and chemical phenomena of polymerization. It consists of a set of differential equations that describe material and energy balances on the reaction mixture. For computational purposes, the model is split into two categories. The first category consists of bulk, suspension, and solution polymerization, while the other describes the emulsion case. In the model development, we shall examine both categories in parallel. Bulk, suspension, and solution polymerizations are characterized by the fact that all of the reaction steps proceed in a single phase. A model for a reactor carrying out such polymerizations would consist of a set of material balances describing the rates of accumulation, inflow, outflow, and disappearance by reaction of the various monomers, initiators, polymers, and other ingredients in the reactor. These polymerizations consist of initiation, propagation, termination, and transfer reactions occurring simultaneously through the full conversion range. Conventional emulsion polymerizations usually occur in three stages and are comprised of more than one phase (reactor head-space; the monomer droplets, which act as a monomer reservoir; the (continuous) aqueous phase, which can act as a locus of polymerization as well as a species transport medium; and the polymer particle phase, the main locus of polymerization). The first of the three common polymerization stages involves the nucleation (birth) of polymer particles. This can occur by either micellar or

homogeneous (coagulative) nucleation. The second stage involves the growth of the particles until the monomer droplets disappear. The third stage begins with the disappearance of the monomer droplets and continues until the end of the reaction. The emulsion polymerization model can be briefly described as follows. First, the initiation can be accomplished via a redox mechanism or via thermal decomposition of an initiator. The fate of radicals (initiator, monomeric, and oligomeric) in the water phase is propagation with dissolved monomers in the water phase, reaction with water-soluble impurities (WSI), termination in the water phase, possible recombination of initiator fragments, reaction with monomer droplets, desorption from polymer particles, reabsorption of desorbed radicals into polymer particles, capture by emulsifier micelles, and capture by polymer particles. The birth of particles can be accomplished by homogeneous (aqueous phase) nucleation, micellar nucleation, and particle coalescence. Once captured by particles (or micelles or droplets), the radicals may propagate, mutually terminate, react with monomer-soluble impurities (MSI), react with chain-transfer agent (CTA), undergo chain transfer to monomer, undergo chain transfer to polymer, and participate in internal and terminal double-bond polymerizations. The average number of radicals per particle is followed by accounting for entry/ absorption of radicals from the water phase, radicalradical termination, radical-MSI termination, and desorption of radicals into the water phase. The partitioning of monomer, monomer-soluble impurities, and CTA into the various phases is another important factor. For both the bulk/suspension/solution and emulsion models, material balances on the various components of the polymerization are used to calculate the conversion, composition, molecular weight, and, in the case of emulsion polymerizations, particle size and number. Other, not necessarily measurable, polymer properties can also be calculated or inferred. Finally, molecular weight averages dependent on termination, branching, and transfer reactions are estimated. The various symbols, subscripts and superscripts, and variables are shown in the Nomenclature section. The units of the variables are also shown therein. 2.1. Initiation. The first step in a polymerization involves the creation of highly reactive free radicals. This is accomplished in the initiation stage. Bulk, suspension, and solution polymerizations involve organicsoluble initiators such as 2,2-azobis(isobutyronitrile) (AIBN). The initiator is decomposed into free radicals by thermal or photochemical (ultraviolet light) means. In emulsion polymerizations, there are two commonly used initiation methods. The first, redox initiation, is used for low-temperature polymerizations, while initiation by thermal decomposition is used for the higher temperature range. Andersen and Proctor (1965) suggested the following mechanism for the redox system persulfate (PS)/sodium formaldehyde sulfoxylate (SFS)/iron (Fe).

S2O82- + Fe2+ 9 SO4- + Fe3+ + SO428 8 Fe3+ + RA 9 Fe2+ + X SO4- + Mj 9 R1,j 8


kpIj k2

k1

(1) (2) (3)

(Symbols not explained in the text are given in the

968 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

detailed Nomenclature section.) In the above mechanism, Mj refers to monomer of type j; S2O82- represents the persulfate initiator, I; RA is the reducing agent (in this case SFS), often referred to as activator; SO4- is the initiator fragment that reacts with monomer, Mj, to produce primary radicals R1,j, i.e., radicals of chain length 1 ending in monomer j. k1 is the rate constant for the oxidation reaction, k2 is the rate constant for the reduction reaction, and kpIj is the propagation rate constant for the addition of monomer j to an initiator fragment. Radicals are generated in the water phase by the reaction between the initiator (PS) and a complex of Fe2+ and ethylenediamine tetrasodium acetate (EDTA) (see eq [REF:eqn:ini1]). The complexation of iron and EDTA reduces the effective concentration of Fe2+ and prevents undesirable side reactions. Thus, the initiator is reduced to a negatively charged free radical, while the Fe2+ is oxidized to Fe3+ (see eq 1). The SFS then reduces Fe3+ to Fe2+ (see eq 2). The free radical, SO4-, reacts with monomer j to form radicals of chain-length unity (see eq 3). Broadhead et al. (1985) used a similar scheme to describe redox initiation for styrene/butadiene (SBR) polymerization. Performing material balances for the initiator (I or PS) and the reducing agent (RA) gives

SO4- + Mj 9 R1,j 8

kpIj

(10)

where kd is the initiator decomposition rate constant. This mechanism results in the following material balance for the moles of initiator:

NI dNI ) FI,in - vout - kdNI dt VT


Finally, the overall rate of initiation, RI, is

(11)

RI ) k1

NFe2+ [I] + 2fkd[I] Vw

(12)

dNI NI k1NINFe2+ ) FI,in - vout dt VT Vw dNRA NRA k2NRANFe3+ ) FRA,in vout dt VT Vw

(4)

(5)

where Ni is the number of moles of component i, Fi,in is the inflow of component i into the reactor in mol min-1, Vw is the volume of the water phase, VT is the total volume of the reaction mixture, and vout is the volumetric flow out of the reactor. The time dependence of all terms involved in the equations is not shown for the sake of brevity. One should note that k1 and k2 are effective rate constants, since other (unknown) elementary reaction steps may be occurring but have been assumed to have a negligible contribution to the overall initiation rate. This will simplify our set of equations. Further simplification occurs when we apply the reactor stationary-state hypothesis to ferrous and ferric ions to provide the ion concentrations in terms of total iron concentration as follows:

where [I] is the initiator concentration and f is the initiator efficiency factor. Equation 12 is implemented into the computer model as shown above and includes the rate of initiation by redox means and thermal means. However, when one of the methods is activated, the term representing the other option becomes negligible. The acceleration of the decomposition of potassium persulfate (KPS) by free sodium dodecyl sulfate emulsifier has been reported by Okubo et al. (1991). Sarkar et al. (1990) reported the acceleration of KPS decomposition due to the addition of vinyl acetate (VAc) monomer, but no emulsifier effect was detected. Considerations such as these may be responsible for some discrepancies between model predictions and experimental data. Often, the initiator efficiency is considered to be constant. However, in a high-viscosity regime the initiator efficiency may decrease significantly (GarciaRubio and Mehta, 1986; Russell et al., 1988b; Zhu et al., 1990a). Since initiator decomposition occurs in the water phase for conventional emulsion polymerizations, it is likely that initiator fragments are not subject to a high-viscosity environment and the initiator efficiency is thus held constant (Adams et al., 1990). One can then argue that, if the initiator efficiency changes, this represents water-soluble impurity effects. For bulk and solution polymerizations, however, the following semiempirical equation is used to describe the changing initiator efficiency when the free volume of the reaction mixture (VF) becomes less than a critical free volume, VFcrif:

f ) fo exp(-C(1/VF - 1/VFcrif))

(13)

NFe2+ )
where

k2NFeNRA k1NI + k2NRA

(6)

NFe ) NFe2+ + NFe3+


and

(7)

dNFe NFe ) FFe,in v dt VT out

(8)

For the case of thermal decomposition of a persulfate initiator, the commonly accepted mechanism is (Sarkar et al., 1988)

S2O82- 9 2SO48

kd

(9)

where fo is the initial initiator efficiency and C is a parameter which modifies the rate of change of the efficiency. The critical free volume, VFcrif, is dependent on temperature and initiator type. The initiator efficiency typically becomes diffusion-controlled at very high conversions (>80 wt %). The calculation of the free volume, VF, will be discussed later. In the model, therefore, eqs 4-7 and 11-13 are used directly to describe the initiation step. 2.2. Water Phase Reactions. The discussion in this section is restricted to emulsion polymerizations. Once the radicals are generated in the water phase, they can then go on to propagate with monomer, react with various other species in the reaction mixture, and nucleate particles. Two approaches for particle nucleation are commonly employed nowadays: homogeneous (coagulative or aqueous phase) nucleation and heterogeneous (micellar) nucleation. Several representative articles dealing with these mechanisms follow: Fitch

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 969

and Tsai (1970, 1971), Hansen and Ugelstad (1978, 1982), Poehlein et al. (1986), Maxwell et al. (1991, 1992b), and Casey et al. (1994). Another interesting, slightly different nucleation case has recently been described by Lepizzera and Hamielec (1994). We now examine the fate of the radical species in the water phase: 1. The radicals may react with monomers in the water phase:

Rn,i + Rm,j 9 Pm+n or Pm + Pn 8

ktw

(18)

8 R + Mj 9 R1,j I
Rn,i + Mj 9 Rn+1,j 8 kpij

kpIj

(14) (15)

where the rate constants of propagation are used to define reactivity ratios as follows:

rij )

kpii kpij

(16)

R represents a radical of chain length n ending in n,i monomer i. Equation 14 represents the reaction between initiator fragments and monomers to form primary radicals. Equation 15 represents propagation. Equation 15 describes the addition of monomer j to a growing radical chain of length n ending in monomer i. The reaction, proceeding with a rate constant of kpij, results in a radical chain of length n + 1 ending in monomer j. For the case of a terpolymerization, there are nine different propagation reactions and six separate reactivity ratios. In principle, kpii in eq 16 may be obtained from homopolymerization data for each monomer type, while the six reactivity ratios may be calculated from the three binary copolymerizations involved in the terpolymerization. Copolymerization is described using four propagation reactions and two reactivity ratios. The propagation reactions shown above represent terminal model kinetics. That is, the reactivity of a radical center is assumed to depend only upon the monomer unit bound in the polymer chain on which it is located. Other alternative models could also be considered and are discussed later. 2. The radicals may react with water-soluble impurities:
8 Rn,i + WSI 9 P(WSI) kzj

where Pm+n is a dead polymer molecule of chain length m + n. Note that in eq 18 termination may occur either by combination of the radical chains or by disproportionation. ktw is often considered to be negligible (Urretabizkaia et al., 1992; Urretabizkaia and Asua, 1994); however, in systems containing highly watersoluble monomers (Sarkar et al., 1988), or depending on the emulsifier concentration (Song and Poehlein, 1988b), water-phase termination may be significant. 4. The initiator fragments may recombine. This phenomenon is taken into account by use of the efficiency factor, f, for the initiation step or by the use of an effective initiator decomposition rate constant. 5. The radicals may be captured by monomer droplets. Most modeling efforts ignore this phenomenon due to the fact that the surface area of the polymer particles and micelles is far greater than that of the monomer droplets. Thus, the likelihood of a radical species entering the monomer droplets is minimal. This assumption can be tested by the use of electron microscopy to determine if any abnormally large particles exist and, if so, how many. Droplet nucleation may occur should extremely high shear rates be used during mixing along with the appropriate emulsifier concentration. Also, this phenomenon usually occurs in the presence of alcohol groups with ionic emulsifiers. The rate of radical capture by monomer droplets can be defined as
Rcmd ) kcmd[RTOT]wdrop[drops]

(19)

where [R ]wdrop is the total concentration of radicals TOT able to enter a micelle, a particle, and/or a droplet (the calculation of this quantity would be similar to that for [R ]wmic and [R ]wpar in eqs 30 and 31, respecTOT TOT tively), and [drops] is the concentration of droplets. In the examples cited later in this paper, the rate constant for the capture of radicals by monomer droplets, kcmd, is set to zero. For the interested reader, the case of monomer droplet polymerization has been described by Ugelstad et al. (1973, 1974), Hansen and Ugelstad (1979), Song and Poehlein (1988a,b), and Fontenot and Schork (1992-93a,b). 6. Radicals may desorb from polymer particles at the following rate:

(17) Fdes )

where P(WSI) is a dead molecule and kzj is the rate constant for reaction of water-soluble impurity, j, with a monomer i-ended radical. The use of kzj as an overall rate constant, regardless of which monomer radical the impurity is reacting with, is a simplification to our model. It is assumed that the rate of reaction of WSIs does not depend on the radical type. Water-soluble impurities may cause an induction period by consuming large amounts of free radicals. Typical examples of such impurities are oxygen and other commonly used monomer inhibitors with a considerable water solubility at the conditions of the polymerization (e.g., hydroquinones). The effects of impurities on polymerization rate and quality are poorly understood, yet they are one of the most important sources of variation in an industrial setting (Huo et al., 1988; Penlidis et al., 1988; Chien and Penlidis, 1994a,b; Dube and Penlidis, 1997). 3. The radicals may terminate upon encountering another radical:

j kdesNpn NAVw

(20)

where Np is the number of polymer particles per liter of water, n is the average number of radicals per j particle, NA is Avogadros number, and Vw is the total volume of water. Several authors have reported expressions for the desorption rate constant, kdes, all of which are based on the same principle. Information may be found in Nomura et al. (1971a), Ugelstad and Hansen (1976), Nomura and Harada (1981), Rawlings and Ray (1988), Mead and Poehlein (1989b), Asua et al. (1989), and Casey et al. (1994). Desorption is a phenomenon restricted to small molecules, usually as a result of chain transfer to monomer or monomer-soluble impurity or chain-transfer agent. However, the use of relatively large CTA molecules (e.g., n-dodecylmercaptan) in many polymerizations would obviate the need to include the chain transfer to CTA in the desorption equation. The desorption rate constant is discussed in detail later.

970 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

7. Desorbed radicals may reabsorb into the polymer particles. Expressions for the reabsorption of previously desorbed radicals back into the particles have been put forth by Poehlein et al. (1986). This phenomenon, while not included in the desorption equation itself, is included in the water-phase radical balances shown later. 8. Radicals may be captured by micelles according to
Rcm ) kcm[RTOT]wmicAm/Vw

method, that of Fitch and Tsai (1970, 1971), involves the use of grouped parameters, thus simplifying the structure of the equations but not necessarily rewarding us with less uncertainty. According to the first method:

RI )

[R]( I

kpIj[Mj]w + kzi[WSI]i) j)1 i)1

Nz

(23)

(21)

where kcm is the rate constant of capture by micelles and Am is the total free micellar area. The radical concentration [R ]wmic is the concentration of radiTOT cals in the water phase which can enter micelles; it does not include those radicals which, due to their size, electric charge, and hydrophilicity, cannot be captured by micelles. The derivation of [R ]wmic is shown TOT later; it is defined in eq 30. Equation 38 gives the rate of particle nucleation by the micellar mechanism. 9. Radicals may be captured by particles at the following rate:
Rcp ) kcp[RTOT]wparAp/Vw

(22)

where kcp is the rate constant for capture by polymer particles and Ap is the total surface area of the polymer particles. The radical concentration [R ]wpar is the TOT concentration of radicals in the water phase which can enter particles. The same arguments explained for micellar capture of radicals in the water phase apply here as well. [R ]wpar is defined later in eq 31. TOT 2.3. Particle Nucleation. The aspect of emulsion polymerization that generates the most discussion and conflict is the debate regarding the nature of particle nucleation (Richards et al., 1989; Dunn, 1992; Hansen, 1992, 1993). In this model, both micellar and homogeneous particle nucleation mechanisms are accounted for. The assumption that particle sizes are monodisperse is also employed in the following discussion; accounting for a distribution of particle sizes is discussed in the next section. Reviews of the polymer particle formation mechanisms may be found in Ugelstad and Hansen (1976) and Hansen and Ugelstad (1982). Micellar and homogeneous nucleation can both play a significant role in particle formation (Fitch and Tsai, 1971; Fitch, 1981). The relative importance of homogeneous nucleation increases with the solubility of monomer in the water phase, while at high emulsifier levels, micellar nucleation dominates due to the high surface areas and rapid radical absorption rates from the water phase. Particleparticle coalescence has been neglected in modeling by most workers. Notable exceptions include Min and Ray (1974), Hansen and Ugelstad (1978), Morbidelli et al. (1983), and Song and Poehlein (1988a,b). It has been difficult to develop a general model for radical absorption into micelles and polymer particles and desorption from polymer particles (Hansen and Ugelstad, 1982; Nomura, 1982). We now proceed to form the expression for homogeneous particle nucleation. In this paper, a combination of two methods, along with modifications to allow for multicomponent polymerizations and inhibitors, was implemented into the model. The first method, originally proposed by Hansen and Ugelstad (1978), is more rigorous but requires knowledge of parameters that are not readily known for all polymer systems. The second

where N is the total number of monomers in the system and Nz is the total number of WSIs in the system. The above expression equates the rate of radical generation to the rate of radical disappearance. In other words, we are invoking a steady-state hypothesis. On the lefthand side of eq 23 we have the overall rate of initiation (see RI of eq 12). The right-hand side of eq 23 shows the disappearance of the initiator radicals by propagation with monomer j in the water phase (see kpIj[Mj]w) and by reaction with water-soluble impurities (see kzi[WSI]i). [R] is the concentration of initiator radicals I in the water phase. [Mj]w is the concentration of monomer j in the water phase. We are making the assumption that initiator radicals will not terminate (this is accounted for by initiator efficiency) and the visualization that radicals of this size will not enter particles nor micelles nor monomer droplets due to electrostatic forces and the hydrophilicity of such radicals. Next, we perform balances on radicals of chain length 1, ending in monomer i:
kpIi[Mi]w[R] ) [R1,i]( I

j)1

kpij[Mj]w +

kzi[WSI]i + i)1
ktw[RTOT]w) (24)

Nz

The above equation now includes termination with other radicals in the water phase (see ktw[R ]w in eq 24). TOT The creation of radicals of chain length 1 by desorption is not included in the above balance. The desorbed radicals are of a different nature compared to primary radicals or oligomers in that they are a result of transfer to monomer and transfer to chain-transfer agent reactions inside the polymer particles. Hence, desorbed radicals have an electric charge different from the other radicals in the water phase and can be recaptured by the particles. In fact, the desorbed radicals will not move far beyond the particle from which they desorbed. That is, it can be argued that the desorbed radicals will not necessarily enter the bulk of the water phase but will stay in the vicinity of the particle surface. However, if there is a large amount of desorption, a significant amount of desorbed radicals may stray from the particles. We now perform balances on radicals of chain length k (2 e k e (jcr/2)), ending in monomer i. jcr/2 is the critical chain length at which oligomers may be captured by micelles, particles, and/or droplets.

j)1

kpji[Rk-1,j])[Mi]w

[Rk,i]( Nz

kpij[Mj]w + j)1
(25)

kzi[WSI]i + ktw[RTOT]w) i)1

Now a balance is performed on radicals of chain length k (jcr/2 + 1 e k e (jcr - 1)), ending in monomer

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 971

i. jcr is the critical chain length at which oligomers will precipitate and form a polymer particle (homogeneous nucleation) given that there is remaining free emulsifier. We also see that all capture mechanisms (micellar, particle, droplet) have now been included (see the terms kcmAm/Vw, kcpAp/Vw, and kcmd[drops], respectively, in eq 26). The inclusion of the capture mechanisms reflects the visualization that the oligomeric radicals have now grown enough so that there are no repulsion barriers (i.e., the charge at the end of the oligomer is no longer strong enough to create a repulsion from micelles, particles, or droplets).

Next, the concentration of radicals in the water phase that may be captured by particles is given by
[RTOT]wpar

i)1

kcpAp/Vw [Rk,i] + kcmAm/Vw + kcpAp/Vw k)(jcr/2)+1 N kcpAp/Vw [Rjcr,i] (31) kcmAm/Vw + kcpAp/Vw + kh i)1
jcr-1

Finally, the concentration of radicals in the water phase that may undergo homogeneous nucleation is defined by
[RTOT]whom )

kpji[Rk-1,j])[Mi]w ) [Rk,i](kpij[Mj]w + j)1 j)1 kzi[WSI]i + ktw[RTOT]w + kcmAm/Vw + kcpAp/Vw + i)1 Nz

kh [Rjcr,i] (32) kcmAm/Vw + kcpAp/Vw + kh i)1


N

kcmd[drops]) (26)
Equations 23-26 represent a reasonable and practical way to handle differing monomer solubilities in water. Our visualization that the capture of radicals by micelles and particles begins at about a length of jcr/2 units has been independently supported by Poehlein (1990), Maxwell et al. (1991, 1992b), and Kshirsagar and Poehlein (1994). jcr is calculated as a weighted function of the instantaneous composition of the polymer formed in the water phase (Fjw):

Am represents the total free micellar surface area, that is, the surface area created by the emulsifier remaining after the coverage of droplets and particles. Am is given by

Am ) ([S]t - [S]CMC)VwSaNA - Ap - Ad Ap ) (Np)1/3(6Vp)2/3

(33) (34)

jcr )

jcrjFjw j)1

(27)

Fjw is described later in eqs 132-134 as Fj. The concentration of radicals of chain length jcr ending in monomer i is

(
[Rjcr,i] )

kpji[Rjcr-1,j])[Mi]w j)1

(28)

[S]t is the total concentration of emulsifier in the reactor, [S]CMC is the critical micelle concentration, Sa is the area occupied by an emulsifier molecule, Ad is the area of monomer droplets, and Ap is the total surface area of polymer particles. As the reaction proceeds, the total surface area of polymer particles quickly becomes very much greater than the total surface area of monomer droplets, so Ad is neglected as a simplification to our model. In eq 34, Vp is the total volume of polymer particles. Equation 34 is based on the assumption that the polymer particles are spherical. Sa (see eq 33) is affected by the polarity of the adsorbing surface and is therefore affected by persulfate initiators (Ali and Zollars, 1985). kh is the homogeneous nucleation rate constant defined by Fitch and Tsai (1971) as

(kcmAm/Vw + kcpAp/Vw + kcmd[drops])

A balance on the total amount of radicals in the water phase (excluding initiator radicals) gives
[RTOT]w )

kh ) kho 1 -

LAp 4Vw

(35)

k)1 [Rk,i] i)1

N jcr-1

L in the above expression represents the critical radical diffusion length and is given by Einsteins diffusion law:

(29) L)

Recall that radicals of chain length jcr/2 + 1 to jcr can be captured by micelles, particles, and droplets. Also, radicals of chain length jcr can undergo homogeneous nucleation. Droplet nucleation is neglected as a simplification to our model. Thus, a balance on the concentration of radicals in the water phase able to be captured (radicals of chain length > jcr/2) yields the following equations. First, the concentration of radicals in the water phase that may be captured by micelles is given by
[RTOT]wmic )

2Dwjcr kpMwsat

1/2

(36)

Mwsat is the saturation concentration of monomer in the water phase. The rate of change of the number of particles is described by the following equation:

dNpVw Vw dNhom dNmic ) FNp,in - Np vout + Vw + Vw dt VT dt dt kFNp2Vw (37)


FNp,in is the inflow of particles (i.e., in the case of a seeded emulsion polymerization), while the second term on the right-hand side of eq 37 (Np(Vw/VT)out) represents the outflow of particles from the reactor. The rates of homogeneous and micellar particle generation are described by the third and fourth terms on the right-hand

i)1

kcmAm/Vw [Rk,i] + kcmAm/Vw + kcpAp/Vw k)(jcr/2)+1 N kcmAm/Vw [Rjcr,i] (30) kcmAm/Vw + kcpAp/Vw + kh i)1
jcr-1

972 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

side of eq 37, respectively. kFNp2Vw represents a crude attempt to take particle coalescence into account (Ugelstad and Hansen, 1976; Hansen and Ugelstad, 1978, 1979). Song and Poehlein (1988a,b), Fontenot and Schork (1992-93a), and Chern and Kuo (1996) discuss particle coagulation in more detail. This is a difficult subject, still under investigation, and hence, especially due to lack of specific data, we decided to neglect the term in further simulations with the model. The rate of micellar nucleation, dNmic/dt, is given by

The volumetric growth rate of polymer in a polymer particle is given by

dVp(t,) dt

)(

MWjRpj(t,)) F (t) j)1


p

V(t,) (42)

V(t,) ) Vp(t,) +
N

Vjmp(t,) j)1

(43)

dNmic ) NAkcm[RTOT]wmic/rmic dt

(38) Rpj(t,) )

fj(t) (

j kpiji(t))[M]p(t) n(t,) i)1 NAV(t,)

As can be seen from eq 30, when the surface area of free micelles, Am, goes to zero, micellar nucleation is halted. The expression for the rate of generation of particles by homogeneous nucleation becomes

(44)

dNhom ) NAkh[RTOT]whomVw dt

(39)

The homogeneous nucleation rate constant, kh, tends toward zero as the area of polymer particles, Ap, increases (see eq 35). This is because there is a higher probability for an oligomer to be captured by a preexisting particle rather than form a new particle by homogeneous nucleation. 2.3.1. Emulsion Particle Size Distribution (PSD) Calculation. If one wishes to relax the assumption of a monodisperse particle size distribution, it is necessary to account for different classes or ages of particles as shown below. In the first instance, it is assumed that statistical broadening can be neglected and, thus, polymer particles born at time with volume V0 will all have the same volume (V(t,)) at some later time t (at least for those particles which have not left the reactor in the case of a CSTR). Calculation of PSD with exact correction for statistical broadening requires the solution of a large number of partial differential equations (Behnken et al., 1963; Sundberg, 1979; Kiparissides and Ponnuswamy, 1981; Rawlings and Ray, 1988; Storti et al., 1989), and this appears to be impractical at the present time except for special cases (the number of polymer particles containing three or more radicals is zero). The consumption rate of monomer in the reactor Rp(t) is given by

where Vjmp(t,) is the volume of monomer j in the particle born at time and i is the mole fraction of radicals in the particles ending in monomer i (see eq 48). Equation 44 is defined by eqs 126 and 86 but with the dependence on t and shown in the variables. The following algebraic relationship may be used to calculate the volumetric growth rate for polymer particles born at times > 0.

dVp(t,) n(t,) dVp(t,0) j ) dt dt n(t,0) j

(45)

In this manner, one can calculate the full particle size distribution. 2.4. Organic Phase Reactions. This section describes reactions in bulk and solution polymerization as well as those reactions occurring in the bulk phase in emulsion polymerization once the radicals have been captured by particles (or micelles or monomer droplets). 1. The radicals may propagate (as shown in eq 15):
8 Rn,i + Mj 9 Rn+1,j kpij

(46)

For multicomponent polymerizations (e.g., in this case, terpolymerization), the overall propagation pseudokinetic rate constant can be defined as (Hamielec et al., 1987; Tobita and Hamielec, 1991; Xie and Hamielec, 1993a,b)

kpo )

kpijifj i)1 j)1

N N

(47) (48)

Rp(t) )

R (t,) V(t,) Np(t,) d 0 p

(40)

1 ) (kp21kp31f12 + kp21kp32f1f2 + kp23kp31f1f3)/

where Rp(t,) is the consumption rate of monomer at time t in polymer particles born at time . Np(t,) d is the number of polymer particles in the reactor at time t which were born at time . The total polymerization rate, Rp(t,) in these particles may be expressed as

2 ) (kp12kp31f1f2 + kp12kp32f22 + kp13kp32f2f3)/ (49) 3 ) 1 - 1 - 2 ) (kp12kp31f12 + kp21kp32f1f2 + kp23kp31f1f3 + kp12kp31f1f2 + kp12kp32f22 + kp13kp32f2f3 + kp12kp23f2f3 + kp13kp21f1f3 + kp13kp23f32) (51) (50)

Rp(t,) )

j kp[M]p(t) n(t,) NAV(t,)

(41)

where kp is defined in eqs 47 and 84, [M]p is the concentration of monomer in the polymer particles, and n is the average number of radicals per particle. j The rates of polymerization for the individual species Rpj(t,) may be found using Rp(t,) and reactivity ratios (see eq 126).

where i is the mole fraction of radicals in the particles ending in monomer i and fj is the mole fraction of monomer j in the particle or bulk phase. The is are calculated as above by invoking the quasi-steady-state assumption (QSSA) for the radicals. The QSSA is not valid at high conversion levels, but the conversion and Mn (number-average molecular weight) results will not h be significantly affected and the Mw (weight-average h

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 973

molecular weight) predictions will be affected only slightly (Achilias and Kiparissides, 1994). The pseudokinetic rate constants for multicomponent polymerization described throughout this paper are in the context of terminal kinetics. Tobita and Hamielec (1991) have derived equivalent expressions for pseudo-kinetic rate constants in the context of the penultimate copolymerization model. The pseudo-kinetic rate constant method is perhaps the only practical way of handling complex multicomponent polymerization modeling. Equations 48-51, as shown above, have been developed for the terpolymerization case. The reduction of eqs 48-51 to copolymerization and homopolymerization and their extensions to higher multicomponent systems should be clear. 2. The radicals may mutually terminate:
Rn,i + Rm,j 9 Pm+n 8 Rn,i + Rm,j 9 Pm + Pn 8 ktd ktc

(52) (53)

Equation 52 represents termination by combination, whereas eq 53 represents termination by disproportionation. The two termination rate constants are defined by the overall termination rate constant, kt, and , the ratio of termination by disproportionation to the overall termination:

) ktd/kt kt ) ktc + ktd

(54) (55)

where P(MSI) is a dead polymer molecule. kfmsi is the rate constant for the reaction of monomer-soluble impurities with radicals ending in monomer j. The use of kfmsi as an overall rate constant, regardless of which monomer the impurity is reacting with, is a simplification to our model. It is assumed that the rate of reaction of MSIs does not depend on the monomer type. In emulsion polymerizations, MSIs are transferred into the particles with monomer(s) during monomer diffusion. In the bulk/solution case, the MSIs are in the same phase as the initiator, thus causing an induction time. The effect of monomer-soluble impurities, usually ignored or at least poorly understood, can be quite pronounced on the overall reaction rate, eventually affecting particle growth (in emulsions), particle nucleation (in emulsions), and molecular weight (Huo et al., 1988; Penlidis et al., 1988; Chien and Penlidis, 1994a,b; Dube and Penlidis, 1997). The impurity effect on particle growth can be described as follows. If a MSI partitions into the growing polymer particles, it scavenges the free radicals. Thus, the particle growth rate is slowed down considerably. This leads to a prolonged particle nucleation stage because the micellar emulsifier, which is used to stabilize the particles, is consumed more slowly. Therefore, the average lifetime of a micelle is extended and more particles are nucleated. Since the rate of polymerization is proportional to the number of particles, a significant increase in the rate may result (Dube and Penlidis, 1997). 4. The radicals may react with a chain-transfer agent such as a mercaptan:
Rn,j + RSH 9 HPn,j + RS 8 kfctaj

Describing an overall termination pseudo-kinetic rate constant in multicomponent polymerizations has been a source of constant debate. In this study, several different methods were attempted, but each method had limitations. The overall termination pseudo-kinetic rate constant was calculated as

(60)

kto )

ktoijij i)1 j)1

N N

(56)

For the bulk/solution case, the cross-termination rate constants were defined as

ktoij ) ktoiFi + ktojFj

(57)

where RSH represents the chain-transfer agent which loses a labile hydrogen to the growing radical chain. kfctaj is the rate constant for transfer to a CTA molecule for the monomer j-ended radical type. The product of the above reaction, RS, continues propagating with monomer(s), thus lowering the average molecular weight. The importance of these reactions is stressed by the findings of Broadhead et al. (1985), who showed that transfer to CTA dominated the SBR polymerization. In the case of multicomponent polymerization, an overall chain transfer to CTA pseudo-kinetic rate constant (kfcta) is defined as

where Fi, the instantaneous polymer composition, is defined later in eqs 132-134. The cross-termination rate constants in the emulsion case were defined as

kfcta )

kfctajj j)1

(61)

ktoij ) (ktoiktoj)1/2

(58)

5. The radicals may undergo chain transfer to monomer:


kfmij Rn,i + Mj 9 Pn,i + M 8 j

There has been little agreement as to how the crosstermination rate constants should be defined. Therefore, both eqs 57 and 58 were implemented for future comparison. An overall was calculated using a weighted average of the homopolymerization s based on the instantaneous polymer composition. 3. The radicals may react with monomer-soluble impurities such as hydroquinone and tert-butylcatechol (TBC), which are commonly added to the fresh monomer by the suppliers due to their radical scavenging properties:
8 Rn,j + MSI 9 P(MSI) kfmsi

(62)

(59)

In this case, the active radical center is transferred to a monomer molecule which may propagate further, thus lowering the overall molecular weight. Chain transfer to monomer is an important reaction. It greatly affects the molecular weight and is a precursor to terminal double-bond reactions and desorption. For example, the transfer reaction to VAc results in a stable radical which reinitiates slowly; thus, desorption from polymer particles is possible (Litt, 1993). As previously with kfcta, an overall pseudo-kinetic rate constant for chain transfer to monomer (kfm) is defined

974 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

as

kfm )

kfmijifj i)1 j)1

N N

k** ) p (63)

k**iFj pij h i)1 j)1

N N

(69)

6. The radicals may undergo chain transfer to polymer:


8 Rn,i + Rm,j 9 Pn,i + Pm,j kfpij

(64)

Transfer to polymer does not necessarily occur at the end of a dead polymer molecule. Thus, the species type at the end of the radical chain is not important. The presence of chain transfer to polymer reactions has been reported for BA and VAc polymerizations (Scott and Senogles, 1970, 1974; Friis et al., 1974; Hamielec, 1981; El-Aasser et al., 1981, 1983; Penlidis et al., 1985a; Dube et al., 1991b; Lovell et al., 1991, 1992). Lovell et al. (1991, 1992) reported that the transfer to polymer reaction for BA occurred by the abstraction of a tertiary hydrogen from a BA unit in the polymer chain. Friis et al. (1974) reported that transfer to polymer and transfer to monomer reactions dominate the VAc polymerization. The chain transfer to polymer occurs at the methyl hydrogen of VAc. The overall pseudo-kinetic rate constant for transfer to polymer (kfp) is

kfp )

h kfpijiFj i)1 j)1

N N

(65)

The cumulative polymer composition, Fj, is defined later h in eq 135. In a batch reactor when there is significant compositional drift, the use of the above equation is not strictly valid and one would have to resort to the cumbersome method of moments for the complete set of chemical equations for polymerization. In many practical circumstances compositional drift is either small or maintained small (batch reactors and reactivity ratio pairs which give small compositional drift, semibatch reactors where compositional drift is controlled to low levels and well-mixed continuous stirred-tank reactors with micromixing where only statistical spreading of composition occurs). 7. The radicals may undergo reactions with internal and terminal double bonds. Once again, an effective rate constant will be used. The rate constant for terminal double-bond reactions is k* and that for p internal double-bond reactions is k**. p
k* p 8 Rn,i + Pm 9 Rm+n,i k** p 8 Rn,i + Pm 9 Rm+n,i

(66) (67)

Terminal double-bond reactions yield trifunctional branch points, whereas internal double-bond reactions yield tetrafunctional branch points. Hence, an overall pseudokinetic rate constant for terminal (k*) and internal p double-bond (k**) reactions are defined as p

k* ) p

k* iFj pij h i)1 j)1

N N

(68)

The importance of the terminal double-bond reaction has been documented for the BA and VAc polymerizations (Scott and Senogles, 1970, 1974; Friis et al., 1974; Hamielec, 1981; El-Aasser et al., 1981, 1983; Penlidis et al., 1985a; Dube et al., 1991b; Lovell et al., 1991, 1992). As in the case for transfer to polymer, the validity of the above pseudo-kinetic rate constant equations comes into question in the presence of strong compositional drift. 2.5. Diffusion-Controlled Rate Constants. When discussing the various reactions taking place in a polymerization, many of the parameters were referred to as rate constants. This is somewhat of a misnomer as these so-called rate constants (e.g., kp, kt) vary with the viscosity of the reaction medium. Thus, in a bulk system, the viscosity increases due to the increase in polymer concentration and, hence, affects the rate of propagation and termination. In the emulsion case, the viscosity in the main locus of polymerization (i.e., the particles) is high from the onset of the reaction due to the high polymer concentration. Thus, the rates of termination and propagation may be diffusion-controlled even at low conversion levels. The general chemical equations for diffusion-controlled termination were expressed earlier in eqs 52 and 53. The termination constants ktc and ktd can be redefined as ktc(n,m) and ktd(n,m), respectively. This illustrates their dependence, in general, on the chain lengths n and m of the polymer radicals, R and Rm,j n,i (see eqs 52 and 53) but not on the polymeric radical type (i and j denote the monomer type on which the radical center is located on the end of the polymer chain). One might expect that two backbone radical centers would mutually terminate at a significantly lower rate. The dependence of this rate on the position of the radical centers in the polymer backbone is not clear and the modeling of this effect is beyond the scope of this investigation. In other words, we limit the treatment of polymer radical/polymer radical termination to linear chain systems where termination rates are controlled by radical centers on chain ends. The termination rate constants for linear chain systems also depend on the mass concentration, molecular weight distribution of the accumulated dead polymer, and polymerization temperature. When most of the dead polymer chains (chains without a radical center) are made via chain transfer to a small molecule (such as chain-transfer agent, monomer, etc.) chain-length dependence of the termination rate constant may not be an important issue (accurate MWDs of the dead polymer may be calculated without a knowledge of this chain-length dependence). The calculation of the total polymer radical concentration and then the rate of polymerization would require at most a single number-average termination constant (ktN). However, when a significant number of h dead polymer chains are made via polymer radical/ polymer radical termination, then to calculate the full MWD or higher molecular weight averages (Mw, Mz, h h etc.), one must account for chain-length dependence when it is operative. When calculating MWD and its averages, chain-length dependence can be accounted for in different ways. One can use an approximate expression to account for the dependence of kt(n,m) on chain

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 975

lengths n and m (kt(n,m) ) ktc(n,m) + ktd(n,m)). Some of the earlier studies to use this approach include Benson and North (1962), Duerksen and Hamielec (1967), Duerksen (1968), and others and more recent studies by Cardenas and ODriscoll (1976), Ito (1980, 1981), Soh and Sundberg (1982a), Coyle et al. (1985), Russell et al. (1992), and OShaughnessy and Yu (1994a,b). These attempts to find a general and effective expression for kt(n,m) based on theoretical and experimental considerations are to be lauded even though to date an expression for kt(n,m) which can be considered effective and useful for commercial polymer reactor modeling is not available. In the meantime, it is recommended that a less general approach using averages of kt(n,m) be used (Boots, 1982; Olaj et al., 1987; Zhu and Hamielec, 1989) for reactor modeling. The use of kt(n,m) averages permits one to calculate the rate of polymerization, Mn, Mw, and possibly Mz of both h h h accumulated dead polymer and dead polymer produced instantaneously. Individual empirical correlations for ktN, ktW, and ktZ versus polymer concentration, temperh h h ature, and other significant variables as shown by Zhu and Hamielec (1989) are required. Methods based on averages have yet to be comprehensively evaluated although Vivald-Lima et al. (1994a) have recently compared the effectiveness of the CCS (Chiu et al., 1983) and MH (Marten and Hamielec, 1979) models. The number-average termination constant is given by

For chemically-controlled termination, kt equals kt, h and these are given by equations such as eq 56. For isothermal polymerization and small compositional drift, kt can be treated as a constant. The instantaneous number-average molecular weight is given by

Mn ) h

MWi(R i)1

Rpi
tc/2

(72)

+ Rtd)

ktN ) h

m)1kt(n,m) nm h n)1
h Rt ) ktNYo2

(70)

and the termination rate of polymeric radicals, Rt, by

(71)

where is the mole fraction of polymeric radicals of n chain length n. The use of ktN to calculate Mw, Mz, and h h h higher molecular weight averages will give estimates that are smaller than the true values when chain-length dependence is significant. For the correct calculation of higher molecular weight averages one should use ktw, h ktz, etc. Examples are given in Zhu and Hamielec h (1989). Anseth et al. (1994a) have made a comprehensive experimental investigation of the effect of volume relaxation on free-radical cross-linking kinetics. This phenomenon may explain observations made by Stickler (1983) of the effect of initiator level on limiting conversions for polymerization of methyl methacrylate in the absence of cross-linking monomers. At higher polymerization rates, the actual volume is greater than the equilibrium volume, permitting higher limiting conversions to be reached. Anseth et al. (1994b) have also recently shown dominance of termination by reaction diffusion in highly cross-linked systems. Recent papers by Vivaldo-Lima et al. (1994b) and Hutchinson (19923) have accounted for diffusion-controlled termination as affected by cross-linking. The validity of the stationary-state hypothesis (SSH) was tested by direct experimentation for the homopolymerization of methyl methacrylate and for the copolymerization of methyl methacrylate and cross-linker ethylene glycol dimethacrylate using ESR for the first time (Zhu et al., 1990b). The SSH was shown to be valid for homopolymerization of methyl methacrylate and for low levels of cross-linking. For high levels of crosslinking, however, the SSH is clearly not valid.

where Rpi is the rate of polymerization of monomer i and Rtc and Rtd are rates of termination by combination and disproportionation, respectively. As mentioned earlier, the modeling to be discussed herein will use kt h exclusively, with chain-length dependence neglected in calculating Mw, Mz, etc. h h Under diffusion-controlled termination a single termination rate constant can be employed to model the rate of multicomponent polymerization and molecular weight development. In other words, the monomer type on which the radical center is located is unimportant. However, when dealing with polymer radicals with long branches, the self-diffusion coefficient of the polymer molecule will depend not only upon chain length or the number of monomer units in the chain but also upon the number of long branches, branch lengths, and their location on the polymer molecule. The position of the radical center (on the chain end or somewhere on the backbone) should also affect the termination rate. Termination under these conditions is clearly very complex, and a good deal of empiricism is required to model termination reactions for polymer radicals with long branches. The self-diffusion coefficient should also depend upon the distribution of monomer types in the polymer chains. This is another complicating factor when compositional drift is important. Diffusion-controlled rate constants are modeled in this paper using the free-volume approach (Marten and Hamielec, 1979; Soh and Sundberg, 1982a-c; Hamielec et al., 1987). In order for the termination reaction to occur, two macroradical chains must approach each other via translational diffusion. Next, the diffusion of the chain segments containing the active centers toward each other occurs. This is termed segmental diffusion. Finally, the termination reaction takes place. Thus, three diffusion-control intervals can be described. In the early stages of an isothermal, batch multicomponent polymerization (typically at conversions less than about 10% in bulk polymerization), where termination might be chemically controlled, compositional drift is usually small and hence a single constant termination constant, kt (kt ) ktc + ktd), may be used to model the rate of polymerization and molecular weight development to the conversion where termination becomes diffusion-controlled. However, when relatively high molecular weight polymers are being produced at low monomer conversions, the termination rate may be controlled by segmental diffusion. Bhattacharya and Hamielec (1986), Jones et al. (1986), and Yaraskavitch et al. (1987) have applied a model for segmental diffusion-controlled termination after Mahabadi and ODriscoll (1977). The rate constant for termination controlled by segmental diffusion is given by

ktseg ) kto(1 + c)

(73)

where kto is given by eq 56 shown earlier, is a parameter dependent on the molecular weight of the polymer radicals as well as solvent quality (monomers

976 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

plus solvent if present), and c is the mass concentration of accumulated polymer in the reaction mixture. As the polymer concentration increases in a good solvent, the coil size of polymer chains decreases, increasing the gradient of concentration of the radical center across the coil and thus the diffusion rate of the radical center. The consequence of this is that the termination rate constant increases with monomer conversion to the point where translational diffusion of the center of mass of the polymer radical coils controls the termination rate and then kt begins to fall very rapidly. The contribution of c is rather small in many cases (Bhattacharya and Hamielec, 1986; Jones et al., 1986; Hamielec et al., 1987; Yaraskavitch et al., 1987) and could be neglected in most reactor calculations. The next interval, translational diffusion control, occurs up to high polymer concentrations (85-90% conversion in bulk). It is in this stage that the gel effect or autoacceleration occurs. Using the free-volume approach, the point at which the reaction becomes translational diffusion-controlled is identified by a temperature-dependent term, K3, defined by Marten and Hamielec (1979, 1982) as

For the modeling of homopolymerizations and copolymerizations, the adjustable parameters have been arbitrarily set equal to 0.5 for m and 1.75 for n (Marten and Hamielec, 1979, 1982; Garcia-Rubio et al., 1985; Bhattacharya and Hamielec, 1986; Jones et al., 1986; Yaraskavitch et al., 1987). Panke (1986) has shown that setting m ) n ) 0.5 gives an equally good fit to rate and molecular weight data for the homopolymerization of methyl methacrylate. These values for m and n also give the correct response when methyl methacrylate is polymerized in a batch reactor with a high molecular weight heel. Finally, in extremely viscous environments, the reaction diffusion-control termination rate constant, ktrd, becomes significant. Two methods of describing this regime were employed in the model. The first (Stickler method) uses an expression by Stickler et al. (1984) and is given by

ktrd )
where

8NAD 1000

(77)

K3 ) Mwcrit exp(A/VFcrit) hm

(74)

where A and m are adjustable parameters but hopefully independent of temperature, radical initiation rate, and monomer and polymer concentrations over wide ranges of these polymerization conditions. This has been proven so for both homo- and copolymerization by Stickler et al. (1984) and others (Bhattacharya and Hamielec, 1986; Jones et al., 1986; Yaraskavitch et al., 1987). Mwcrit is the accumulated weight-average moh lecular weight at the monomer conversion at which the termination rate is translational diffusion-controlled. Mwcrit = Mwcrit, the instantaneous weight-average moh lecular weight, and thus there is a direct connection with the size of the polymer radicals. VFcrit is the critical free volume corresponding to that conversion. K3 was found to have an Arrhenius temperature dependence for both homo- and copolymerization (Stickler et al., 1984; Bhattacharya and Hamielec, 1986; Jones et al., 1986; Yaraskavitch et al., 1987). The frequency factor had a small dependence on chain composition for the p-methylstyrene/acrylonitrile system (Yaraskavitch et al., 1987). In multicomponent polymerizations, the free volume is given by

( )
6Vm NA

1/3

(78)

D)

nslo2 k [M] 6 p

(79)

and NA is Avogadros number, is the reaction radius, D is the reaction diffusion coefficient, Vm is the molar volume, ns is the number of monomer units in one polymer chain segment, and lo is the length of the monomer unit. The second method (RNG method) involves what is termed residual termination. Russell et al. (1988a) defined upper and lower bounds for the residual termination rate parameter as

4 ktres,min ) kp[M]a2 3
and

(80)

8 ktres,max ) kp[M]a3jc1/2 3

(81)

VF )

(0.025 + Ri(T - Tgi))V i)1

Vi
T

(75)

where i represents the monomer, polymer, and solvent. Ri is the difference in the thermal expansion coefficients for species i above and below its glass transition temperature, Tgi. T is the polymerization temperature, Vi is the volume of each species, and VT is the total volume of the reaction mixture. The translational diffusion-controlled termination rate parameter, kT, is given by

where a is the root-mean-square end-to-end distance per square root of the number of monomer units, is the Lennard-Jones diameter, and jc is the entanglement spacing of pure polymer, measured in monomer units. When attempting to model multicomponent polymerizations, options to use residual monomer mole fraction weighted averages or overall values for a, , and jc were made available. The lower limits of the RNG method and the Stickler method give similar results. The upper and lower bounds are handled in the model with the following expression:

ktrd ) ktres,minx + ktres,max(1 - x)

(82)

kT ) ktcrit

( ) ( (
Mwcrit h Mw h
n

1 1 exp -A VF VFcrit

))

(76)

where ktcrit is the value of kt when eq 74 is satisfied. n is an adjustable parameter.

where x is the conversion. As will be discussed later, the reaction diffusion (or residual termination) concept may not be perfectly correct. Recent ideas regarding trapped radicals may have to be incorporated into future modeling efforts (Zhu et al., 1990a).

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 977

The point at which the reaction changes from translational diffusion control to reaction diffusion control can be somewhat ambiguous. Thus, the following overall kt is defined to circumvent this problem:

kt ) ktseg + kT + ktrd

(83)

Equation 83 has been previously employed to handle the various diffusion-control equations for kt by Marten and Hamielec (1979) and Soh and Sundberg (1982ac). Vivaldo-Lima et al. (1994a) showed that this approach (which they termed as a serial approach) resulted in as good, if not better, prediction of various polymerization data compared to the parallel approach such as that used by Russell (1994). The parallel approach involves summing the reciprocals of each kt expression from each diffusion-control interval. This is discussed further by Gao and Penlidis (1996). As will be discussed later, the prediction of the termination rate is still under investigation. Recent publications by Zhu et al. (1990a), Tobita and Hamielec (1991), Achilias and Kiparissides (1992), Maxwell and Russell (1993), Buback et al. (1994), Russell (1994), and Tobita (1994a,b) have shown some insight into solving this problem while at the same time demonstrating a wide variety of approaches. For free-radical polymerizations below the glass transition temperature of the polymer being synthesized, reaction mechanisms at very high monomer conversions can be very complex due to the decrease in the diffusion coefficients of small molecules: primary radicals and monomer. These decreases in mobility will have one or more of the following consequences: (1) decrease of the initiation efficiency; (2) the propagation reaction becomes diffusion-controlled; (3) radical pair formation (Zhu and Hamielec, 1989). An experimental observation for such polymerizations is the reduction in conversion rate to almost zero, even though appreciable concentrations of monomer and initiator exist in the polymerization mixture. Most polymerization models permit the propagation constant to fall with monomer conversion while fixing the initiator efficiency at a constant value. This is in effect equivalent to letting the product f1/2kp fall with monomer conversion to fit the rate of polymerization. There are practical reasons for doing this because it is very difficult to measure f and kp separately without the use of electron spin resonance spectroscopy (ESR). The use of data for polymerization rate and for molecular weights of accumulated polymer can, in principle, permit one to estimate f and kp separately as has been done before (Hui and Hamielec, 1968; Duerksen and Hamielec, 1968). However, for the bulk polymerization of methyl methacrylate with the very pronounced diffusion-controlled termination followed by a glassy effect and the associated broad MWD with both low and high molecular weight tails on the distribution, accurate molecular weight measurements are extremely difficult if not impossible to obtain. ESR is the preferred method of measuring both f and kp at high monomer conversions for bulk polymerization of monomers such as methyl methacrylate. Zhu and Hamielec (1989) did ESR/rate of polymerization measurements and found that f and kp fall significantly with conversion at about thesame monomer conversion, while Ballard et al. (1986) observed that for the same monomer (methyl methacrylate) but with emulsion polymerization, kp became diffusion-controlled at higher monomer conversions. This interesting result may be due to the swelling of surface layers of polymer particles by water in

emulsion polymerization. Sulfate groups dragged into the interior of polymer particles may be responsible for this. In the event that f falls sooner than kp, the dead polymer produced in the time interval between the fall in f and kp would have very high molecular weights. This should affect high-order molecular weight averages. This, however, has not been confirmed experimentally for MMA bulk polymerization. Molecular weights are difficult to measure as already mentioned, and there is the possibility that chain transfer to monomer might intervene to put a cap on molecular weights generated. For a discussion of phenomena such as radical pair formation and heterogeneous effects during glassy-state transition, one should refer to the original paper by Zhu and Hamielec (1989). In order to model diffusion-controlled propagation, i.e., when the polymerization temperature is less than the glass transition temperature of the polymer being synthesized, the propagation rate constants are then given by

kpij ) kpijo exp -B

( (

1 1 VF VFpcrit

))

(84)

where i is the radical type where the active center is located, j is the monomer type being added to the polymer chain, B is an adjustable parameter which should depend on monomer molecule type, and VFpcrit is the critical free volume where the propagation reaction of monomer j adding to a polymer chain ending in monomer i becomes diffusion-controlled. One might expect these critical free volumes to depend on polymer radical type as well as monomer type with propagation rates which are greater, becoming diffusion-controlled at lower monomer conversions. It is also reasonable to expect that diffusion-controlled propagation rate constants for the same monomer are equal (kpij ) kpji). This can be used to reduce the number of adjustable parameters. When dealing with multicomponent polymerizations, the diffusion-control expressions shown above were simplified somewhat by taking weighted averages of many of the parameters. The parameter from the segmental diffusion-controlled termination rate expression (see eq 73) was averaged by the cumulative polymer composition. A weighted average of the parameter A from eq 74 by the instantaneous polymer composition (Fi) was included as an option in the simulation package. As well, the sum of the K3 parameters of each monomer was included as an option. The option for the VFpcrit parameter in the diffusion-controlled propagation rate expression (see eq 84) was a weighted average based on the residual monomer mole fraction. Alternatively, overall values for A, K3, and VFpcrit were input. Although several parameters were mentioned above as being adjustable, only the latter three (A, K3, VFpcrit) were adjusted to fit the data. As will be discussed later, they were not adjusted for each experiment. 2.6. Free-Radical Concentrations. In order to calculate the reaction rates, one must also know the concentration of free radicals in the reactor. For the bulk, suspension, and solution case, one may use the overall concentration of free radicals in the reactor, but for the emulsion case, one requires an average number of radicals per particle. 2.6.1. Bulk/Suspension/Solution Case. The concentration of free radicals in the bulk, suspension, and solution cases is given by

978 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

( Yo )

i)1

Nz

kzi2[MSI]i2 + 4ktRI)1/2 2kt

i)1

Nz

kzi[MSI]i (85)
and

kcpAp([RTOT]wpar + [RTOT]des)VpNA2

R)

Np2kt

(90)

Equation 85 describes the initiation of radicals by initiator decomposition and the consumption of free radicals by reaction with impurities. 2.6.2. Emulsion Case. In the emulsion case the overall concentration of free radicals in the particle phase is

m)

(kdes + kfmsi[MSI]p)VpNA Npkt

(91)

Yo )

nNp j VpNA

(86)

where n is the average number of radicals per polymer j particle. There have been several methods proposed to estimate n (Stockmayer, 1957; OToole, 1965; Ugelstad j et al., 1967; Ballard et al., 1981a; Huo et al., 1988; Li and Brooks, 1993). One can, in principle, write a balance around a polymer particle assuming a steady state, where the disappearance of radicals from the particle is equal to the entry of radicals into the particle. The phenomena to be accounted for in the balance are radical entry (absorption, capture) from the aqueous phase, radical-radical termination in the particle, reaction of the radical with monomer-soluble impurities in the particle, and desorption of radicals into the water phase. The balance is performed for Nn, the number of particles with n radicals:
kcp([RTOT]wpar + [RTOT]des)[Nn] + kt(n) (n - 1)[Nn]/vp + (kdes + kfmsi[MSI]p)n[Nn] ) kcp([RTOT]wpar + [RTOT]des)[Nn-1] + kt(n + 2)(n + 1)[Nn+2]/vp + (kdes + kfmsi[MSI]p)(n + 1)[Nn+1] (87)

where [R ]des is the concentration of desorbed radiTOT cals in the water phase given by
[RTOT]des )

FdesVw kcpAp

(88)

vp in eq 87 is the average volume of a particle or Vp/Np. (Also note that Nn and Np in eqs 87, 90, and 91 represent the number of particles, not the number of particles per liter of water.) The first term (on the left-hand side) in eq 87 represents the capture of free radicals (including desorbed radicals) from the water phase (see also eq 31). The second term describes termination in the particles, while the third term represents radical desorption and reaction with monomer-soluble impurities. The solution of the above recursive equation was originally proposed as a modified Bessel function; Ugelstad et al. (1967) proposed the following partial fraction expansion as an approximation:

n) j m+ m+1+

R 2R 2R m+2+ 2R m + 3 + ...

(89)

where

Equation 89 requires approximately 10 levels of fractions to obtain a converged value for the average number of radicals per particle. More recently, Li and Brooks (1993) presented a more universal model (does not assume a steady state) for estimating n. An investigation of the Li and Brooks j (1993) approach is recommended in future modeling efforts. In the computer implementation, we decided to adopt the method of Ugelstad et al. (1967) due to its simple form and reduced amount of computational effort. 2.7. Radical Desorption. We will now look at the rate of free-radical desorption from the polymer particles (restricted to the emulsion polymerization case). In the most general case, desorption occurs when a small radical traverses from the particle to the aqueous phase. The desorbed radical is usually assumed to be no more than 1 monomer unit in length. The formation of these monomeric radicals in the particles is a result of chain transfer to monomer, to impurity, or to chain-transfer agent. In our case, radicals formed as a result of chain transfer to chain-transfer agent will be assumed not to desorb since typical CTAs are relatively large and insoluble molecules (e.g., n-dodecylmercaptan) (Nomura et al., 1982). It is commonly accepted that desorbed radicals will not move far beyond the particle from which they desorbed. That is, it can be argued that the desorbed radicals will not necessarily enter the bulk of the water phase but will rather stay in the vicinity of the particle surface. However, if there is a large amount of desorption, a significant amount of desorbed radicals may stray from the particles. Thus, while the following model equations are rigorous, there are certain underlying assumptions that should always be checked with specific process data, in order to explain possible discrepancies. Expressions describing desorption rate constants continue to be developed and reported in the literature. For large particles, desorption may be negligible; however, in the early stages of a reaction (i.e., when the particles are small) or when the rate of absorption of radicals by the particles is slow, desorption may be important. In the following equations, any reactions with monomer soluble impurities were assumed to yield nonreactive products. Several groups have developed expressions for the desorption rate constant. Their expressions attempted to give the desorption rate constant as a function of particle diameter. Mead and Poehlein (1989b) derived particle-size-independent expressions for kdes. They also rederived the equations put forth by other research groups while incorporating particle-size independence. The notion of a nonuniform distribution of radicals in the particles is applicable to large particles and in cases where chain-transfer reactions do not dominate (Chern and Poehlein, 1990). It is due to the hydrophilicity of end groups (initiator fragments) on the oligomers being captured. de la Cal et al. (1990) also discussed the nonuniform distribution of radicals. As a simplification

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 979

to our model, particle-size-independent expressions were not used. Recently, Fontenot and Schork (1992-93a) used the model developed by Nomura et al. (1976) and incorporated ideas by Mead and Poehlein (1989a,b) to account for nonuniform monomeric radical concentrations in particles. Fontenot and Schork (1992-93a) were dealing with large particles and droplet nucleation. The first desorption expression that we will examine is that of Ugelstad and Hansen (1976):

kdes )

12Dw kfm kp (a + D /D )d 2 w p p

(92)

The parameter (see eq 94) can be described as a lumped coefficient or, alternatively, as the ratio of film mass-transfer resistance to overall mass-transfer resistance (Nomura and Harada, 1981). At the start of the polymerization, Dw and Dp are of the same order of magnitude because the particles are saturated with monomer. Thus, assumes a value close to unity. As the conversion and, hence, the viscosity inside of the particles increases, Dp decreases markedly and, hence, kdes decreases. An empirical model of the decrease in Dp was used in the model (Friis and Hamielec, 1977):

In eq 92, kfm is the transfer to monomer rate constant, kp is the rate constant for reinitiation of oligomeric radicals from desorbed monomeric radicals, Dw is the diffusivity of monomer radicals in the aqueous phase, a is the partition coefficient for monomer radicals between the aqueous and particle phases, Dp is the diffusivity of monomeric radicals in the particles, and dp is the particle diameter. Harada et al. (1971) and Nomura et al. (1971a) have theoretically derived an expression for the desorption rate coefficient, kdes, using a stochastic approach and have successfully applied it to vinyl acetate emulsion polymerization. Their result is the following:

Dp ) Dpo

1 ((1 - - x ) 0.19x

+ 0.0017x

(97)

j kdes ) n +

kp[M]pmddp2 12Dw

)(
-1

kfm kfcta[CTA]p + + kp kp[M]p j RI(1 - n) Npkp[M]pn j

Previous efforts in the literature to simulate emulsion polymerizations (Harada et al., 1971; Nomura et al., 1971b; Friis and Nyhagen, 1973; Nomura et al., 1976) made use of eq 96 and assumed an average constant value for throughout the conversion range. Thus, in our case, eqs 94 and 95 were used. This approach is identical to the one used by Penlidis (1986) to model vinyl acetate and vinyl chloride emulsion polymerizations. Furthermore, in order to accommodate the multicomponent aspect of the polymerization, the parameters Dw and md were calculated as weighted functions of the mole fraction of monomer in the particles. Rawlings and Ray (1988) presented an expression that included chain transfer to CTA. However, their use of an overall chain-transfer rate constant would enable one to modify other groups equations similarly.

kp[M]p (93)

In eq 93, md is the partition coefficient for monomeric radicals (same as Ugelstad and Hansens a and equivalent to 1/Kjwp in eq 110), dp is the average particle diameter, Dw is the diffusivity of monomer radicals in the aqueous phase, kfm is the transfer to monomer rate constant, and

ktr 3Dm kp kdes ) 3DmMwt + r2 dmkp


where

(98)

) 1+

6Dw mdDp

-1

(94)
and

Dm )

D p Dw mdDp + Dw

(99)

where Dp is the diffusivity of monomeric radicals in the particles. In eq 93, the term RI(1 - n)/Npkp[M]pn j j represents desorption of initiator radicals. In our model development, it was assumed that initiator radicals would not be captured by the particles due to their hydrophilic nature and electric charge; hence, this term may be safely neglected. As discussed above, radicals formed as a result of chain transfer to CTA will not desorb due to the size of the CTA assumed in this case. Thus, kfcta[CTA]p/kp[M]p in eq 93 can be neglected. Equation 93 can then be written as

ktr ) ktrm 1 +

ktrt[CTA]p ktrm[M]p

(100)

j kdes ) n +

kp[M]pmddp2 12Dw

-1

ktrt is the chain transfer to CTA rate constant, [CTA]p is the concentration of CTA in the particle, ktrm is the chain transfer to monomer rate constant, dm is the monomer density, is the monomer volume fraction in the particle, Mwt is the monomer molecular weight, and r is the particle radius. Asua et al. (1989) offered the following expression:

kfm[M]p

(95)

Nomura and Harada (1981) derived exactly the same expression as in eq 95 using a deterministic approach. If n is much less than 0.1 during the whole range of j conversions (which is not always the case), then the following expression may be used:

Ko kdes ) kfm[M]p Ko + kp[M]p


where

(101)

12Dw kfm kdes ) m d 2 kp


d p

( )

Ko ) (96)
and

12Dw/mddp2 1 + 2Dw/mdDp

(102)

980 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

kp[M]w + ktw[R]w kp[M]w + ktw[R]w + kaNTw/NA w

(103)

ktw is the rate constant for termination in the water phase, [M]w is the concentration of monomer in the water phase, ka is the radical absorption (by particles) rate coefficient, NT is the total number of particles per unit volume of aqueous phase, w is the volume fracw tion of water in the continuous phase, and NA is Avogadros number. Finally, the particle-size-independent expression derived by Mead and Poehlein (1989b) is shown below:

interaction parameters would permit one to make reasonable estimates of these concentrations. There are several methods available to define partitioning of species between several phases: a rigorous thermodynamic approach, an experimental approach (based on xc, the conversion at which the monomer droplets disappear), an empirical approach, and the use of partition coefficients. Several groups in the literature employ the following equation (Dougherty, 1986):

[M]psat )

1 - xc a2Fm Mwt 1 - xc(1 - Fm/Fp)

(106)

kdes )

kfm 12Dw(/6)2/3 kp

a+

Dw[sinh()(1 + 3/2) - 3 cosh()/]

kp[M]pR2[cosh()/ - sinh()/2] (104)

where

)r

kp[M]p Dp

where a2 is an adjustable parameter that has a value less than 1 for a water-soluble monomer and is equal to 1 otherwise. Fm and Fp are the densities of monomer and polymer, respectively. xc is the critical conversion at which the monomer droplets disappear. After the conversion has reached xc, one can replace xc in the above equation with the conversion level x. The thermodynamic approach presented by Guillot (1985) is as follows:

1/2

(105)

i ) o,i + RT(ln i + (1 - m)j + ijj2) (107)


Equation 107 represents the chemical potential of species i in the monomer droplets. i is the volume fraction of component i, m is a coefficient related to volume characteristics, and ij is the Flory-Huggins interaction parameter. In the water phase:

R, in the above equations, is the radius of a monomerswollen particle and r is the radial position within the monomer-swollen particle. Both the Nomura group and the Poehlein group present expressions for the desorption rate constant for copolymerization. The extension to the copolymer case from the homopolymerization equations shown above is quite straightforward. If one so desires, one may use a weighted average of two homopolymer equations. Also, one should check whether one of the comonomers is insoluble. This could lead to further simplifications and modifications to the copolymer equations. After careful evaluation of the methods presented above, we performed some tests with representative numbers and found very little difference between the various methods. Hence, we have employed the expression for kdes of Nomura and Harada (1981) (see eq 96). As an example, consider the monomers in the BA/ MMA/VAc system. VAc is known to exhibit significant radical desorption (Nomura et al., 1971a; Friis and Nyhagen, 1973; Penlidis, 1986), while MMA desorbs to a lesser extent (Ballard et al., 1981b) and the desorption for BA is considered negligible (Mallya and Plamthottam, 1989). Urretabizkaia et al. (1992) and Urretabizkaia and Asua (1994) reported that desorption was negligible for the seeded semibatch emulsion polymerization of BA/MMA/VAc. Considering the dominance of the acrylic monomers at lower conversions (<60 wt %, see Dube and Penlidis, 1995a-c), it is likely that desorption will not play a role in the BA/MMA/VAc emulsion polymerizations. 2.8. Species Concentrations in the Different Phases. We shall now look at defining the concentration of monomers and of various other components that may partition between different phases (i.e., water phase, monomer droplet phase, particle phase). Under equilibrium conditions, the concentrations of monomers in the particles are determined by the balance between the gain in interfacial free energy caused by an increase in surface area upon swelling and the loss in free energy caused by dissolving the monomer in the polymer. A knowledge of interfacial tension and the Flory-Huggins

w ) w + RT(ln i + (1 - mi/mw)w + miiw2) i o,i (108)


and for the particle phase a similar expression can be written. Asua et al. (1989), Dimitratos et al. (1989), Mead and Poehlein (1989b), Urquiola et al. (1991), Noel et al. (1994), and Schoonbrood et al. (1994) have also used this thermodynamic approach successfully. It is evident from eqs 107 and 108 that several parameters (e.g., the Flory-Huggins interaction parameter, ij; the volume coefficients, mi and mw) must be known. The uncertainties around these parameters make use of the thermodynamic approach less practical compared to the one described below, which is based on defining partition coefficients. Noel et al. (1994) recently published a simplification to the thermodynamic approach which was extended for the case of multiple monomers by Schoonbrood et al. (1994). While the thermodynamic approach is rigorous, partitioning may also be a function of emulsifier concentration and monomer feed (fi) (Guillot, 1985; Emelie et al., 1991). According to the empirical approach, one tries to correlate the points where the monomer droplets disappear with the weight fraction of monomer in the initial charge. This approach involves a good deal of experimentation and will, of course, be practical only when narrow ranges of reactor operation are of interest. Finally, the method of partition coefficients based on that proposed by Omi et al. (1987) was adopted in the current model with ideas from Maxwell et al. (1992a) and Urretabizkaia and Asua (1994). Recent updates of these ideas are also available (Armitage et al., 1994; Gugliotta et al., 1995a). This general and valid approach for modeling uses partition coefficients between the monomer/water phases and the water/polymer phases combined with the overall material balances on

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 981

the monomers, the polymer, and the water. Partition coefficients for monomer j between the monomer droplet phase and the water phase (Kjmw) and for monomer j between the water phase and the particle phase (Kjwp) are defined as

Kjmw ) Kjwp )

Vjmd/Vmd Vjmw/Vaq Vjmw/Vaq Vjmp/Vp

(109)

various phases can then be calculated. The above method has a rigorous thermodynamic basis as shown in Mead and Poehlein (1988); at the same time, the approach is quite practical and amenable to experimental confirmations. In order to obtain the concentration of the monomersoluble impurities and the chain-transfer agent in the particles, partition coefficients were also used as shown below:

(110)

[MSI]p ) [CTA]p )
where

Nmsi Vp + VmKmsi Ncta Vp + VmKcta

(119)

where Vjmd, Vjmw, and Vjmp are the volumes of monomer j in the droplets, aqueous phase, and particles, respectively. Vmd is the volume of the monomer droplets, Vaq is the volume of the aqueous phase, and Vp is the volume of particles. The total volume of monomer j, Vmj, is given by

(120)

Vmj ) Vjmd + Vjmw + Vjmp

(111)

Kmsi ) KMSIwpKMSImw Kcta ) KCTAwpKCTAmw

(121) (122)

where Vjmd and Vjmw are derived from eqs 109 and 110, respectively, as follows:

Vjmd )

VmdVjmpKjmwKjwp Vp VaqVjmpKjwp Vp

(112)

Vjmw )

(113)

Combining eqs 111-113 gives

Vjmp )

Vmj VmdKjmwKjwp VaqKjwp 1+ + Vp Vp

(114)

The algorithm used to calculate the partitioning of the monomers into the different phases is as follows. Initially, when there are no particles present, the volumes of monomer j in the droplet and water phases are calculated by guessing values for Vaq (using monomer solubilities) and Vmd and iterating over the following equations:

Vjmd ) Vmj 1 +

Vaq VmdKjmw

(115) (116) (117)

Equations 119 and 120 both assume that monomersoluble impurities and chain-transfer agents are not considerably soluble in water. Nmsi and Ncta represent the total number of moles of MSI and CTA in the reactor, respectively. 2.9. Material and Population Balances: Productivity and Quality. We now present the various material balances for all of the species involved (save the initiating system which was presented earlier) and the rate equations to which they are related. For the emulsion case, the terms with subscripts p and w represent the particle and water phases, respectively. In the bulk, suspension and solution cases, recalling that the reaction steps proceed in only one phase, the terms with the subscript w should be ignored, while those with subscript p represent the bulk reaction phase. Note that terms for semibatch and continuous reactors which allow for the inflow and outflow of the various species have been included here. In this section, the inflow of species i is designated by Fi,in and vout represents the volumetric flow out of the reactor. The balances for the monomers comprise the inflow and outflow of monomers and their depletion by polymerization in the organic (or particle) phase as well as in the aqueous phase (for the emulsion case):

Vjmw ) Vmj - Vjmd Vmd )

dNmj Nj ) Fmj,in - vout - RpjpVp - RpjwVw (123) dt VT


The inflow of monomer increases the rate of accumulation of the number of moles of monomer j (dNmj/dt) in the reactor. The outflow (second term on the right-hand side of eq 123) and the reaction of monomer (third and fourth terms on the right-hand side of eq 123) result in a decrease in dNmj/dt. The reaction takes place in the particles, hence the term RpjpVp in eq 123, and in the water phase, hence the term RpjwVw in the same equation. Similarly, the balances for the monomers bound in the polymer involve accumulation of bound monomer by polymerization in both the particles and the water phase. In a batch reactor, where there is no inflow or outflow of a polymer from the reactor, the total amount of polymer formed and its composition can be obtained from the monomer material balances directly by considering the net change in moles of each monomer j.

Vjmd j)1 Vjmw j)1


N

Vaq ) Vw

(118)

When particles are generated, the volume of particles, Vp, is calculated using the differential equations shown later in eqs 141 and 142. The volume of the aqueous phase, Vaq, is calculated from the moles of monomer in the aqueous phase (see eq 125). The volume of the monomer droplet phase, Vmd, is calculated by the difference between the total monomer in the reactor and the sum of the monomer in the particle and water phases. Then, eq 114 is calculated followed by eqs 112 and 113. The concentrations of the monomers in the

982 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

However, in a semibatch or continuous-flow reactor with inflow and outflow of polymer and monomer, additional balances are needed for the moles of each monomer j in the reactor that are bound as polymer, i.e.

F1 ) F3 f1(f1r23r32 + f2r23r31 + f3r21r32)(f1r12r13 + f2r13 + f3r12) f3(f1r12r23 + f2r13r21 + f3r12r21)(f1r32 + f2r31 + f3r31r32) (130) F2 ) F3 f2(f1r13r32 + f2r13r31 + f3r12r31)(f1r23 + f2r21r23 + f3r21) f3(f1r12r23 + f2r13r21 + f3r12r21)(f1r32 + f2r31 + f3r31r32) (131)
After some algebraic manipulation, one may obtain

dNpolj Npolj ) Fpj,in v + RpjpVp + RpjwVw (124) dt VT out


The moles of monomer j bound in the polymer (Npolj) are a result of polymerization in the particles (RpjpVp) and in the water phase (RpjwVw) as well as inflow (Fpj,in) and outflow ((Npolj/VT)vout). Polymerization in the monomer droplets is neglected for the reasons presented earlier. A separate balance on the number of moles of monomer j in the water phase is used to calculate the volume of monomers in the water phase. During the interval when monomer droplets exist, the number of moles of monomers in the water phase is held constant from the initial values based on the partitioning equations shown earlier (see eq 116). When the droplets are depleted, the monomer in the water phase is calculated by:

F1 )

num(F1/F3) - 2num(F2/F3) - 2den num(F1/F3) - num(F2/F3) - den num(F2/F3) num(F1/F3) - num(F2/F3) - den den num(F1/F3) - num(F2/F3) - den

(132)

F2 ) F3 )

(133)

dNmjw ) -RpjwVw dt
The polymerization rates are defined as

(125)

(134)

Rpjp ) Yo[M]pfj

kpiji i)1 kpijiw i)1


N

(126)

Rpjw ) [RTOT]w[M]wfjw

(127)

where num(Fi/Fj) refers to the numerator of the equations for Fi/Fj (see eqs 130 and 131) and den refers to their denominator. Equations 132-134 may be reduced to the homo- or copolymerization case by setting the reactivity ratios, which include the missing monomer, equal to 1. The cumulative polymer composition (Fj) is calculated h as

where is the total concentration of radicals in the water phase (see eq 29). From the above balances, a number of variables of interest can be calculated directly. For example, the total mass conversion of monomers to polymer can be calculated by

[R ]w TOT

Fj ) h

Npolj

Npolj j)1
The instantaneous polymer composition is calculated mainly for its use as a weighting function for the various parameters in the model and is calculated for two separate cases. One is based on the polymer being formed in the particles and the other on that being formed in the water phase. On the other hand, the cumulative polymer composition takes into account the polymer formed in both phases, i.e., the total polymer formed in the reactor. The cumulative polymer composition includes both water phase and particle phase polymerizations in order to account for the contribution to the composition due to the capture of oligomeric radicals from the water phase by the particles. A balance is now performed on the solvent, which is in the emulsion case, water, comprising of inflow and outflow:

(135)

x)

(NpoljMWj) j)1 ((Nmj + Npolj)MWj) j)1


N

(128)

The instantaneous polymer composition (Fj) in the particles can now be calculated as:

Fj )

Rpjp

Rpjp j)1
Expressions for the instantaneous polymer composition based solely on the reactivity ratios (defined earlier in eq 16) and the monomer feed mole fractions were derived from the Alfrey-Goldfinger equations for terpolymers (Alfrey and Goldfinger, 1944). The AlfreyGoldfinger equations are shown below:

(129)

dNs NS ) Fs,in v dt VT out

(136)

Similarly, a balance for the emulsifiers results in

Nem dNem ) Fem,in v dt VT out

(137)

Note that the emulsifier can be located in four areas:

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 983

as micelles, stabilizing monomer droplets, and stabilizing particles and dissolved in the aqueous phase. Now, balances for the water-soluble (z1, z2) and monomer-soluble impurities (MSI) are performed:

dNzi Nzi ) Fzi,in v - kziNzi[RTOT]w dt VT out dNmsi Nmsi ) Fmsi,in v - kfmsi[MSI]pYoVp dt VT out

to the density differences between the monomers and the polymer. What results is a net decrease in the total volume of particles. The particle diameter of the swollen particles, under the monodispersed approximation, is defined as

(138) dp ) (139)

( )
Ap Np 6Vpp Np

1/2

(143)

Water-soluble impurities such as oxygen and hydroquinones can enter the reaction via various feed streams and consume radicals in the water phase. Monomersoluble impurities can arise from the feed streams and recycle streams. A balance for the consumption of chain-transfer agent (CTA) is

The particle diameter of the unswollen particles (the quantity measured using a disc centrifuge), under the monodispersed approximation, is defined as

dp )

( )

1/3

(144)

Ncta dNcta ) Fcta,in v - kfcta[CTA]pYoVp (140) dt VT out


The volume of polymer particles is defined in two separate regimes: when monomer droplets exist and when there are no longer any droplets. The balances respectively are

The volume of the monomer droplets can be calculated by the difference between the volume of the organic phase and the sum of the volume of polymer particles and the volume of monomer in the aqueous phase. The balance for the volume of the organic phase is

dVo ) dt

j)1

Fmj,inMWmj Fmj
N

dVp ) dt

MWmj(RpjpVp + RpjwVw) j)1


pFp

MWmj(RpjpVp + RpjwVw) F j)1


(141)

1 mj

1 Fp

(145)

dVp ) dt

MWmjRpjwVw j)1
Fp

MWmjRpjpVp(1/Fmj - 1/Fp) j)1

(142)

where p is the volume fraction of polymer in the particles, Fmj is the density of monomer j, and Fp is the density of the polymer. Fp is calculated using the individual polymer densities weighted by the mole fractions of each polymer in the particles. For the case of seeded emulsion polymerizations or a series of constant stirred-tank reactors (CSTRs), a term describing the inflow of polymer particles can be added to the equations. Note that one may be required to include an outflow term to describe the flow of polymer out of the reactor. Since the outflow term, vout, includes both the organic (monomer and polymer) and aqueous phases, it should be multiplied by the percentage of polymer in the reactor. In any case, one must satisfy the mass balances. The first term in eq 141 describes the increasing volume of the particles due to the saturation of the particles by monomer. In other words, as polymerization within the particles proceeds, monomer is consumed and replaced by monomer from the droplets in amounts required to maintain a constant monomer concentration in the particles. RpjwVw takes into account the growth of the particles by capture of oligomeric radicals formed in the aqueous phase. Equation 141 is active while the monomer droplets exist. Equation 142 represents the stage of the emulsion polymerization when the monomer droplets disappear as a separate phase. At this stage, the volume of the particles can increase only by capture of radicals from the water phase. This growth is counteracted by shrinkage due

The volume of the organic phase changes due to the inflow of monomer and polymer, the outflow of monomer and polymer, and volume shrinkage during the polymerization. Vo corresponds to the total volume in a bulk polymerization. Recall that vout can be included in the equation to describe the outflow of monomer and polymer. 2.10. Molecular Weight, Long-Chain Branching, and Cross-Linking Development. 2.10.1. Linear Copolymer Chains. With linear copolymer chains, polymer molecules once formed are inert and do not take part in reactions such as transfer to polymer and reaction with internal double bonds bound in polymer chains, and, therefore, the method of instantaneous molecular weight distribution can profitably be employed for homo-, co-, and, in general, multicomponent polymerizations. This approach, shown below, was first introduced for binary copolymerization by Stockmayer (1945), who derived the instantaneous bivariate distribution for linear copolymer chains, assuming that both monomer types have the same molecular weight. A correction factor to account for differing monomer molecular weights was recently derived by Tacx et al. (1988) and is given by eq 210 shown later.

W(r,y) ) ( + )( + /2( + )r)r exp(-( + )r(1/22)1/2 exp(-y2/22) (146)


where W(r,y) dr dy is the weight fraction of copolymer of chain length in the range r to r + dr and composition deviation in the range y to y + dy for copolymer produced instantaneously (in the time interval t to t + dt in a batch reactor).

y ) y1 - F1 2 ) F1(1 - F1)K/r

(147) (148)

984 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

K ) (1 - 4F1(1 - F1)(1 - r1r2))

(149)

where y1 is the mole fraction of monomer 1 in the copolymer and F1 is the overall mole fraction of monomer 1 in the copolymer and was given earlier in eq 129 and

ktdYo kp[M]p

+ Cfm +

Cfcta[CTA]p [M]p ) ktcYo kp[M]p +

Cfmsi[MSI]p [M]p

(150)

(151)

The pseudo-kinetic rate constants for multicomponent polymerization were given earlier by eqs 61 (for transfer to CTA), 63 (for transfer to monomer), 83 (for diffusioncontrolled termination), and 84 (for diffusion-controlled propagation). The instantaneous molecular weight distribution of the copolymer is given by

W(r) )

-W(r,y) dy ) ( + )( + /2( + )r)r

exp(-( + )r) (152)


For long polymer chains, one can assume with small error that all chains produced instantaneously have the same composition F1, F2, .... These overall compositions can be found using eq 129. The bivariate distribution for polymer accumulated in a batch or semibatch reactor may be found by integration of eq 146 as follows:

0tW(r,y) RpW(t) V(t) dt W(r,y) ) h 0tRpW(t) V(t) dt


RpW(t) )

(153)

RpiMWi i)1

(154)

chain-length-dependent termination is significant may be found in Zhu and Hamielec (1989). 2.10.2. Polymers with Long Branches. The method of instantaneous molecular weight distribution which is such a powerful method for the calculation of molecular weight distribution of linear polymer chains must be modified when long-chain branching reactions such as transfer to polymer and reaction with internal or terminal double bonds are significant. In this case, dead polymer chains are no longer inert and can react with polymeric radicals (and other radical types) and now the instantaneous molecular weight distribution is no longer a permanent quantity but changes with time as its chains react. One must keep track of these chains in order to predict the MWD of the accumulated linear chains. In principle, this can be done for chains having 1, 2, or more long branches (Soares and Hamielec, 1996c). Alternatively, one can use Monte Carlo simulations to calculate the full MWD for chains with long branches as illustrated by Tobita (1995a,b). Teymour and Campbell (1994) developed a novel method coined numerical fractionation to calculate the full MWD of sol before and after the gelation point without the normal discontinuity that plagued earlier methods using the method of moments. This method is presently being comprehensively evaluated by various research groups, and we await the results of these studies. Of course, when the full MWD is not essential and average molecular weights will do, it is recommended that the method of moments be used. The use of the pseudo-kinetic rate constant method for copolymerization with long-chain branching has been illustrated by Xie and Hamielec (1993b,c). Using the method of moments, the leading moments of the distribution (Q0, Q1, Q2) can be calculated as follows. Invoking the stationary-state hypothesis for radicals, one can readily derive the following moment equations (Ray, 1972; Broadhead et al., 1985; Hamielec et al., 1987; Xie and Hamielec, 1993a,b).

Instantaneous number- and weight-average chain lengths are given by

dVpQ0 CkQ1 KQ0 ) + /2 k [M]pYoVp (159) dt [M]p [M]p p dVpQ1 ) dt

rN ) rW )

1 + /2 2 + 3 ( + )2

(155) (156)

Number- and weight-average chain lengths of the accumulated polymer in batch or semibatch reactors are given by

1 + Cfm +

Cfcta[CTA]p [M]p -

Cfmsi[MSI]p [M]p

kp[M]pYoVp (160)

j rN )

0tRpW(t) V(t) dt
(1/rN)RpW(t) V(t) dt 0
t

dVpQ2 KQ1 + CkQ2 bracket ) factor + 2 1 + + dt denom [M]p


where

(157)

(bracket) )k [M] Y V denom


2 p p o

(161)

0trWRpW(t) V(t) dt j rW ) 0tRpW(t) V(t) dt

(158) ) ktdYo kp[M]p + Cfm + Cfcta[CTA]p [M]p ) ktcYo kp[M]p + Cfmsi[MSI]p [M]p (162)

Tobita and Hamielec (1991) generalized Stockmayers bivariate distribution to account for chain-length dependence of polymer radical/polymer radical termination. Expressions equivalent to eqs 155 and 156 when

(163)

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 985

factor ) 1 + Cfm +

Cfcta[CTA]p [M]p +

Cfmsi[MSI]p [M]p

(164)

denom ) + +

CfpQ1 [M]p KQ1 + + [M]p Cfmsi[MSI]p [M]p

(165)

assumed that transfer to monomer and reaction with internal double bonds control molecular weight development, with an insignificant number of polymer chains produced by termination reactions. In addition, it is assumed that, in the batch reactor, the volume of the reacting mixture is constant. Equation 161 thus reduces to

bracket ) 1 + Cfm +

(Cfp + Ck)Q2 [M]p Cfcta[CTA]p [M]p

dQ2 2 ) (1 + Q2Ck/[M]p)2 kp[M]pYo dt Cfm (166)

(177)

Introducing monomer conversion, eq 177 can be rearranged to give

Cfm, Cfcta, and Cfmsi are the constants for transfer to monomer, to chain-transfer agent, and to monomersoluble impurities, respectively. They are defined as

dQ2 h Q2 h ) 2 1 + dx 1-x
where

(178)

Cfm ) kfm/kp Cfcta ) kfcta/kp Cfmsi ) kfmsi/kp

(167) (168) (169)

Q2 ) k**Q2/kp[M]p h p
and

(179)

Cfp, Ck, and K are the constants for transfer to polymer, internal double-bond, and terminal double-bond reactions, respectively. They are defined as

) k**/kfm p
Equation 178 can be solved analytically to give

(180)

Cfp ) kfp/kp Ck ) k**/kp p K ) k*/kp p

(170) (171) (172)

Q2 ) h

4(1 - x)(1 - (1 - x) (1 + 4 -

1-x 1 + 8) 1 - xc

[( )

1+8

1+8

-1

(181)

The accumulated number- and weight-average molecular weights are then defined as

0.2 0.3 0.4 0.5 1.0 5.0 : 0 0.1 xc: 1 0.943 0.835 0.778 0.647 0.577 0.370 0.094
where xc is the monomer conversion at the point of gelation. The solution (eq 181) applies for a homopolymerization of a divinyl monomer whose double bonds have equal reactivity. For the copolymerization of a divinyl monomer with a vinyl monomer, one must redefine as follows:

Mw ) Q1/Q0Mweff h Mw ) Q2/Q1Mweff h

(173) (174)

The effective molecular weight, Mweff, is calculated by summing the individual monomer molecular weights and weighting them by the cumulative polymer composition. To calculate instantaneous number- and weightaverage chain lengths, one can simply divide eq 160 by eq 159 and then eq 161 by eq 160, respectively. Setting Cfp, Ck, and K to zero should then give the average chain lengths for linear polymer chains (see eqs 155 and 156). The branching frequencies for both trifunctional and tetrafunctional branching are calculated as follows:

) F1k**/kfm p

(182)

dVpQ0BN3 ) kpYoVp(CfpQ1 + KQ0) dt dVpQ0BN4 ) kpYoVpCkQ1 dt

(175) (176)

2.10.3. Cross-Linking and Gelation. Reaction with internal double bonds may rapidly lead to a threedimensional network and the onset of gelation. In the context of the present kinetic models, the onset of gelation occurs when Q2 (or the weight-average molecular weight) goes to infinity. In other words, the first cross-linked gel particle of infinite molecular weight forms in the polymerizing mixture. To illustrate the use of eq 161 to calculate the onset of gel formation, a limiting form of this equation will be used. It is

where F1 is the mole fraction of divinyl monomer in the copolymer chains. The physical interpretation of these results is straightforward. The more active the double bonds bound in the copolymer chains (larger k**) and the greater the p number of these double bonds (larger F1), onset of gelation occurs at lower monomer conversions. The more transfer to a small molecule such as monomer or chain transfer agent (larger kfm), the further in conversion is the onset of gelation delayed. Models which can be used to calculate the quantity of gel and details of gel structure are under development (Tobita and Hamielec, 1989, 1992; Tobita, 1992, 1993a,b; Charmot and Guillot, 1992; Teymour and Campbell, 1994). 2.11. Energy Balance Equations. Two energy balances, one around the reactor contents and the other around the reactor cooling jacket, will enable one to investigate nonisothermal operation, variable heattransfer coefficients, reactor runaways, and temperature controller design). For heat transfer through the reactor walls (reactor jacket) and with the jacket full of perfectly mixed heat-transfer fluid, the two heat balance equations may be written as follows for a well-mixed vessel:

986 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

Reactor Balance d(

MiCpi(T - Tref)) j i

) Fi,inCpi(Tin - Tref) dt i Fi,outCpi(T - Tref) + Rpi(-Hpi)Vp - UAj(T Tj) - QR (183)

Cooling Jacket Balance dTj (FjwCpw(Tj,in - Tj) + UAj(T - Tj) - Qj) ) (184) dt FwVjCpw
Controller Equation. Reactor temperature can be controlled by manipulating the temperature of the cooling jacket water feed. The temperature adjustment of the cooling jacket water feed is assumed to be instantaneous. A proportional-integral controller can be simulated as follows:

Tj,in ) Kce (

Kc I

0te dt

(185)

where the error e is the difference between the desired and actual reactor temperature.

e ) T - Tset

(186)

The integral term of the controller can be simulated with the following differential equation

dI/dt ) T - Tset
where

(187)

I)

0te dt

(188)

The control action is implemented through the following equation:

Tj,in ) Tc + Tj,in

(189)

where Tj,in represents the control action and Tc a steady-state or bias term. The controller gain, Kc, and integral time, I, can be tuned using a linearized process model and stability analysis. U and Aj are the overall heat-transfer coefficient and jacket heat-transfer area, respectively. QR and Qj are uncontrolled heat losses for reactor and jacket. 3. Model Development: Olefin Polymerization with Homogeneous and Heterogeneous Ziegler-Natta Catalysts Polyolefins are very important commodity polymers. Polyethylene and polypropylene are today the major tonnage plastic materials worldwide, accounting for 44% of all U.S. plastic sales in 1988 and reaching a capacity of about 45 million tons in 1990 (Elias, 1992; Whiteley et al., 1992). Most industrial processes for the production of polyolefins utilize heterogeneous Ziegler-Natta catalysts. Conventional soluble Ziegler-Natta catalysts have not found widespread industrial applications, mainly because of insufficient catalytic stability and stereochemical control. This picture, however, will likely change

in the near future with the advent of metallocene-based catalyst systems. Metallocene catalysts with aluminoxane and other cocatalysts are able to produce polyolefins at a very high productivity with a degree of microstructural control not possible to achieve using conventional Ziegler-Natta catalysts (Hamielec and Soares, 1996; Huang and Rempel, 1995; Soares and Hamielec, 1995a). Ziegler-Natta catalysts are formed by a transition metal salt of metals of groups IV-VIII and a metal alkyl of a base metal of groups I-III (known as cocatalyst or activator). However, not all combinations are equally efficient and can be used for all monomers. From the industrial point of view, most heterogeneous catalysts are based on titanium salts and aluminum alkyls. Most types of heterogeneous Ziegler-Natta catalysts have a common intriguing characteristic: they yield polymer with broad molecular weight distribution (MWD) and, in the case of copolymerization, broad chemical composition distribution (CCD). There is now a general agreement that heterogeneous Ziegler-Natta catalysts possess more than one type of active site, each one with distinct ratios of chain transfer to propagation rates, comonomer reactivity ratios, and stereoselectivities. Each site type makes polymer chains that have different average chain lengths, comonomer compositions, comonomer sequence lengths, and, in the case of asymmetric monomers, different degrees of stereoregularity. Consequently, heterogeneous Ziegler-Natta catalysts produce a mixture, at the molecular level, of polymer chains having dissimilar average properties. These dissimilar average properties are reflected in broad MWDs and CCDs. Additionally, intraparticle heat- and masstransfer resistances during the polymerization may somewhat broaden these distributions even further. On the other hand, polyolefins made with most soluble Ziegler-Natta catalysts, including several metallocene systems, have narrow MWD, and copolymers also have narrow CCD. This behavior supports the multiple-site-type hypothesis for heterogeneous catalysts. Soluble Ziegler-Natta catalysts consist of reasonably well-defined, single catalytic species, probably not subject to significant heat- and mass-transfer resistances during polymerization. The complexity of MWDs and CCDs of polyolefins made with Ziegler-Natta catalysts constitutes a challenging problem for polymer quality control. Most properties of polyolefins are routinely measured only as average values. Measurements of melt flow index (as an estimator for a molecular weight average), melt flow index ratio (as an estimator of polydispersity), and bulk density (as an estimator of copolymer composition or degree of short-chain branching) are common practice in industry. However, macroscopic properties of polymers, in general, and polyolefins, in particular, cannot be uniquely determined by average values, since polymers that have some average properties in common can possess other properties that differ markedly. Even a knowledge of the full molecular weight distribution for polypropylene may not be sufficient for many practical applications, due to differences in stereoregularity of the chains. In the same way, determining average compositions of copolymers or average degrees of branching will not entirely define the polymer in question. The whole distribution of composition in addition to chain length is necessary to accomplish this task. This issue becomes very complex with polyolefins made with heterogeneous

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 987

Figure 1. Levels of mathematical modeling.

Ziegler-Natta catalysts because polymers with broad and sometimes multimodal MWDs and CCDs can be produced. The concern about the breadth of MWD and CCD of polyolefins is far from academic. Those distributions affect the final mechanical and rheological properties of polyolefins and ultimately determine their applications. Polyethylenes with broad MWD are easier to process because of greater flowability in the molten state at high shear rate, while polyethylenes with narrow MWD have greater dimensional stability, higher impact resistance, greater toughness at low temperatures, and higher resistance to environmental stress cracking. For polypropylene, narrow MWD is required for rapid molding of products with good mechanical properties. Narrow MWD and especially narrow CCD are required for the production of elastomers. Polymers with broad CCD might have some chains containing long sequences of one monomer type, increasing the crystallinity of the polymer, which is highly undesirable for elastomers. 3.1. Mathematical Models. It is convenient to classify mathematical models for polymerization processes in three levels: microscale, mesoscale, and macroscale (Ray, 1988). Microscale models define the kinetics of polymerization and the types of active sites on the catalyst. Mesoscale models define interparticle and intraparticle mass- and heat-transfer resistances in the polymer particle. Macroscale models describe the macroscopic behavior of the polymerization reactor, such as imperfect mixing, residence time distribution, gasliquid mass transfer, and removal of heat of polymerization. The final application of the model determines the degree of complexity required in each modeling level. Figure 1 shows schematically these three levels of mathematical modeling. Knowledge of the kinetics of polymerization is evidently fundamental for the mathematical modeling of Ziegler-Natta polymerization reactors. The kinetics of polymerization with heterogeneous and homogeneous Ziegler-Natta catalysts has been the subject of detailed

study and will not be covered in this review. Good reviews on polymerization kinetics are available in the literature for both heterogeneous (Keii, 1972; Cooper, 1976; Bohm, 1978; Boor, 1979; Tait and Watkins, 1989) and homogeneous (Hamielec and Soares, 1996; Huang and Rempel, 1995; Soares and Hamielec, 1995a) Ziegler-Natta catalysts. Mathematical models that can predict MWD and CCD of polymer chains will be reviewed in the next sections. 3.1.1. Heterogeneous Catalysts. Heterogeneous Ziegler-Natta catalysts are porous secondary particles, formed by loosely aggregated primary particles (Tait, 1989, Noristi et al., 1994). During polymerization, the growing polymer chains fragment these secondary particles, forming an expanding particle containing primary particles and living and dead polymer chains (Figure 2). This catalyst fragmentation mechanism has been documented for several types of heterogeneous Ziegler-Natta catalysts (Hock, 1966; Buls and Higgins, 1970; Wilchinsky et al., 1973; Boor, 1979; Kakugo et al., 1989; Noristi et al., 1994). One of its consequences is the well-known replication phenomenon: the particlesize distribution of the polymer particles at the end of batch or semibatch polymerization closely approximates the particle-size distribution of the catalyst at the beginning of polymerization (Simonazzi et al., 1991). Good replication is supposed to occur when there is an adequate balance between the mechanical strength of the particle and catalytic activity. Replication factors of 40-50 (ratio of average polymer particle diameter to average catalyst particle diameter) are obtained with third generation Ziegler-Natta catalysts (Galli and Haylock, 1992). For the case of continuous reactors, residence time distribution can significantly affect this particle size distribution (Soares and Hamielec, 1995d; Choi et al., 1994). Based on this well-known experimental evidence, some researchers advocate that, due to diffusion resistances, catalyst fragments in different radial positions are exposed to different concentrations of monomer and

988 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

Figure 2. Fragmentation of heterogeneous Ziegler-Natta catalyst (secondary) particles during polymerization. The growing polymer chains break the secondary catalyst particles into primary particles, forming an expanding particle made of catalyst fragments surrounded by dead and living polymer chains. This behavior is responsible for the replication phenomenon observed in heterogeneous ZieglerNatta catalysis.

Figure 3. Generic representation of physical models. T ) temperature; [M] ) monomer concentration; %M1 ) percent of comonomer type 1 in the polymer; Mn ) number-average molecular weight. In physical models, intraparticle mass- and heat-transfer resistances are responsible for MWD and CCD broadening.

chain-transfer agent (generally hydrogen) and consequently produce polymer with chain-length averages that differ spatially inside the polymeric particle. For copolymerization, monomers with different effective diffusivities and reactivities may be responsible for spatial compositional heterogeneity in the polymeric particle. In addition, if there is appreciable heattransfer resistance, hot spots can occur inside the polymer particle, altering reaction rates and further broadening MWD and CCD (Figure 3). Models based on this approach are generally called physical models (Schmeal and Street, 1971, 1972; Singh and Merrill, 1971; Crabtree et al., 1973; Nagel et al., 1980; Taylor et al., 1983; Floyd et al., 1986a-c; Galvan and Tirrell, 1986a; Hutchinson and Ray, 1988; Skomorokhov et al., 1989; Sarkar and Gupta, 1991, Soares and Hamielec, 1995b).

A second category of mathematical models neglects heat- and mass-transfer resistances in the polymer particle. In these models, multiple types of catalytic sites are responsible for the production of polymer with broad MWDs and CCDs ( chemical models). Each site type has its own kinetic constants and produces polymer with often very different MWDs, CCDs, and stereoregularities, as shown in Figure 4 for a three-site-type catalyst (Tait and Wang, 1988; de Carvalho et al., 1989; Rincon-Rubio et al., 1990; McAuley et al., 1990; Lorenzini et al., 1991; Vela-Estrada and Hamielec, 1994; Lee et al., 1994; Soares and Hamielec, 1995b). There is vast experimental evidence supporting the existence of multiple site types on heterogeneous Ziegler-Natta catalysts (Cozewith and VerStrate, 1971; Boor, 1979; Keii, 1982; Zucchini and Cecchin, 1983; Keii et al., 1984; Chien et al., 1985; Usami et al., 1986; Spitz, 1987; Cheng and Kakugo, 1991; Soares and Hamielec, 1996a,b). The effects of transfer resistances and of multiple site types on MWD and CCD broadening are evidently not exclusive; the heterogeneity caused by the presence of multiple site types can be increased by transfer resistances during the polymerization. Some models combining these two effects have been proposed recently and seem to be highly successful in predicting the qualitative behavior of polymerization with Ziegler-Natta catalysts. These models will be called hybrid models since they combine the two previous modeling strategies (Galvan and Tirrell, 1986b; Floyd et al., 1988; Ray, 1988; Hutchinson et al., 1992; Sau and Gupta, 1993; Soares and Hamielec, 1995b). Physical Models. Different mathematical models have been proposed to describe interparticle and intraparticle mass and heat transfer during polymerization with heterogeneous Ziegler-Natta catalysts. Some are very simplified pictures of the polymerization process and are only useful as a reference point for comparison with more sophisticated models. Among those, the most

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 989

Figure 4. Generic representation of chemical models. 1, 2, and 3 indicate sites of different types; open and shaded circles indicate different types of monomers. In chemical models, multiplicity of active site types is responsible for MWD and CCD broadening.

Figure 5. Schematic representation of the solid core model and of the polymeric core model.

representative for this review are the solid core model and the polymeric core model (Figure 5). The solid core model does not model the breakup of the catalyst particle. The polymer is considered to grow around a solid catalyst core containing all active sites on its surface. This model using a single type of catalyst site cannot predict broad MWDs (Schmeal and Street, 1971, 1972; Nagel et al., 1980). This result should not come as a surprise since the polymerization takes place only at the surface of the catalyst. For a given polymerization time, the monomer concentration is constant at the surface of the catalyst and therefore all polymer chains are produced with the same average properties. Only if the concentration of monomer at the surface of the catalysts changed significantly in time (because of increasing transfer resistance caused by the growth of polymer around the catalytic particle) would the solid core model predict significant broadening of MWD. Additionally, the solid core model is clearly in contradiction with experimental data on catalyst breakup. Despite of these limitations, Brockmeier and Rogan (1976) adjusted the solid core model to data collected in a

semibatch reactor and used this model to predict the behavior of a continuous reactor. The predictions were accurate for polymer yield, but no molecular weight results were shown. With the polymeric core model, polymer grows around a nonexpanding polymeric core formed by polymer and catalyst particles (Figure 5). Although this model is an improvement over the first one, it still is not able to account for broad MWDs (Schmeal and Street, 1971, 1972; Singh and Merrill, 1971). It shows, however, the importance of a core of polymer-catalyst fragments in the explanation of broad MWD. The models that best represent the polymerization in this category of single-site-type, transfer-controlled models are the so-called expansion models. Expansion models consider the fragmentation of the catalyst particle and the formation of an expanding particle of polymer and catalyst fragments. Growing polymer chains and catalyst fragments form a continuum in the polymeric flow model (Schmeal and Street, 1971, 1972; Singh and Merrill, 1971; Galvan and Tirrell, 1986a). Diffusion of reagents as well as heat transfer occurs in this polymeric particle. If the reaction is diffusion-controlled, the radial profiles of monomer and chain-transfer agent in the particle may cause MWD and CCD broadening. The effective mass- and heat-transfer parameters in the polymeric particle have to be estimated to use this model (Figure 6). The polymeric flow model can predict significant broadening of MWD using reasonable physical parameters. The multigrain model (Crabtree et al., 1973; Nagel et al., 1980) considers two levels of mass- and heat-transfer resistances. The polymeric particle ( macroparticle or secondary particle) is formed by an agglomerate of microparticles or primary particles. Each microparticle consists of a fragment of the original catalyst particle, with all active sites on its surface, surrounded by dead

990 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

described by the well-known diffusion-reaction equation in spherical coordinates:

Ms Ms 1 ) 2 Dsrs2 - rv t rs r rs
s

(190)

Ms (r )0,t) ) 0 rs s Ds
Figure 6. Schematic representation of the polymeric flow model and of the multigrain model.

(191)

Ms (r )Rs,t) ) ks(Mb - Ms) rs s Ms(rs,t)0) ) Ms0

(192) (193)

and living polymer (the solid core model describes the behavior of the microparticles). Monomer diffuses through the pores of the macroparticle and through the layer of polymer surrounding the catalyst fragment in the microparticle. Polymerization occurs on the surface of the catalyst fragment. Electron microscopy studies confirm the formation of primary and secondary structures in polymerizations with Ziegler-Natta catalysts (Hock, 1966; Buls and Higgins, 1970; Kakugo et al., 1989; Noristi et al., 1994). It is necessary to estimate the number of microparticles and the transfer parameters in the microparticle and macroparticle to use the multigrain model (Figure 6). The multigrain model is probably the most studied expansion model for polymerization of olefins with heterogeneous Ziegler-Natta catalysts. It has been extensively used to study heat- and mass-transfer resistances for the homopolymerization of ethylene and propylene in slurry and gas-phase reactors. Some important conclusions obtained with the multigrain model are as follows: (1) In most cases, intraparticle temperature gradients are negligible. (2) Concentration gradients in the macroparticles are likely to be more important in slurry reactors. For gas-phase reactors, these gradients may be significant in the microparticles. (3) Interparticle mass- and heat-transfer resistances are negligible except for high activity catalysts. (4) For slurry reactors, gas-liquid mass-transfer resistances are generally negligible for sparged reactors but can be important for unsparged reactors (Nagel et al., 1980; Taylor et al., 1983; Floyd et al., 1986a-c). Most massand heat-transfer resistances are only significant at early stages of polymerization as suggested by detailed mathematical modeling of catalyst particle fragmentation (Laurence and Chiovetta, 1983; Ferrero and Chiovetta, 1987a,b, 1991). Hutchinson and Ray (1988) used the multigrain model to analyze the influence of monomer adsorption on the surface of the microparticle. In the multigrain model, before diffusing into the microparticle, adsorption of the monomer on the outer surface of the microparticle has to take place. Rate enhancement in copolymerization and differences between slurry and vapor-phase reactions were explained based on differences of solubilities in each case, since more crystalline polymer layers have lower monomer solubility and effective diffusivity. Hutchinson et al. (1992) further expanded the multigrain model to describe copolymerization and particle morphology. Modeling equations for intraparticle radial profiles of monomer concentration and temperature for all expansion models are very similar. The radial profile of monomer concentration in the secondary particle is

where Ds ) effective diffusivity of monomer in secondary particle (cm2 s-1), ks ) mass-transfer coefficient in an external film (cm s-1), Mb ) monomer concentration in the reactor (mol cm-3), Ms ) monomer concentration in the secondary particle (mol cm-3), Ms0 ) initial monomer concentration in the secondary particle (mol cm-3), rv ) volumetric rate of polymerization is the secondary particle (mol cm-3 s-1), rs ) radial position in the secondary particle (cm), Rs ) radius of the secondary particle (cm), and t ) polymerization time (s). Notice that rv is the average rate of polymerization at a given radial position in the secondary particle. In the multigrain model, the polymerization is supposed to only take place on the surface of the catalyst fragment in the primary particle. Mathematically, it is introduced as a boundary condition for the mass balance equation of the primary particles. The radial profile of monomer concentration in the primary particles is given by the equations for the solid core model:

Mp Mp 1 ) 2 Dr2 t rp p p rp r
p

(194)

4Rc2Dp

Mp 4 (r )Rc,t) ) Rc3rc rp p 3

(195) (196) (197)

Mp(rp)Rp,t) ) Meq e Ms Mp(rp,t)0) ) Mp0

where Dp ) effective diffusivity of monomer in the primary particle (cm2 s-1), Meq ) equilibrium concentration of monomer in the interface between the primary and secondary particles (mol cm-3), Mp ) monomer concentration in the primary particle (mol cm-3), Mp0 ) initial monomer concentration in the primary particle (mol cm-3), rc ) rate of polymerization on the surface of the catalyst fragment (mol cm-3 s-1), Rc ) radius of the catalyst fragment in the primary particle (cm), rp ) radial position in the primary particle (cm), and Rp ) radius of the primary particle (cm). The rate of polymerization on the surface of the catalyst fragment is generally given by

rc ) kpC*Mc

(198)

where kp ) average rate constant for monomer propagation (cm3 mol-1 s-1), C* ) concentration of active sites on the catalyst surface (mol cm-3), and Mc ) monomer concentration on the catalyst surface (mol cm-3).

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 991

The effective diffusivity in the secondary particle can be estimated as usually done for heterogeneous catalytic reactions (Froment and Bischoff, 1990), using the value of monomer diffusivity in the reaction medium, Db:

Ds ) Db /

(199)

where and are the porosity and tortuosity of the secondary particle, respectively. For the primary particles, one has to correct the diffusivity of monomer in amorphous polymer, Da, according to the equation:

Dp ) Da/R

(200)

where R and are correction factors to account for polymer crystallinity and chain immobilization of the amorphous fraction (Floyd et al., 1986b). Analogous equations can be used for the radial temperature profiles. For the secondary particle:

FpCp

Ts Ts 1 ) 2 kers2 + (-Hp)rv (201) t rs r rs


s

Ts (r )0,t) ) 0 rs s ke Ts (t )Rs,t) ) h(Tb - Ts) rs s Ts(rs,t)0) ) Ts0

(202)

(203) (204)

where Cp ) heat capacity of the secondary particle (cal mol-1 K-1), h ) film heat-transfer coefficient (cal cm-2 s-1 K-1), Hp ) heat of polymerization (cal mol-1), ke ) effective heat conductivity in the secondary particle (cal cm-1 s-1 K-1), Tb ) reactor temperature (K), Ts ) temperature in the secondary particle (K), Ts0 ) initial temperature in the secondary particle (K), and Fp ) density of the polymer particle (mol cm-3). Similarly, for the primary particles the temperature profile is given by

FpCp

Tp Tp 1 ) 2 kerp2 t rp r rp
p

(205)

-4Rc2ke

Tp 4 (r )Rc,t) ) (-Hp) Rc3rc (206) rp p 3 Tp(rp)Rp,t) ) Ts Tp(rp,t)0) ) Tp0 (207) (208)

where Tp0 is the initial temperature in the primary particle. Equations for intraparticle monomer concentration and temperature radial profiles for the polymeric flow and related models are similar to the ones for the secondary particles of the multigrain model. The reaction term in eqs 190 and 201 are calculated using the average concentrations of monomer, M, and catalytic h sites, C*, in the secondary particle: h

solved simultaneously with the intraparticle mass- and heat-transfer equations for each radial position in order to calculate the molecular weight and composition averages in the polymeric particle. The existence of yet another level of mass transfer is suggested by Skomorokhov et al. (1989). In their model, microparticles (described by the solid core model) agglomerate to form subparticles. The polymer particle (macroparticle) results from the agglomeration of these subparticles. Growing polymer rapidly fills the pores of the subparticles, and therefore the monomer diffusion in the subparticle is in the order of magnitude of diffusion in pure polymer. This model evidently predicts higher diffusion resistances than that predicted by the multigrain model, since the effective diffusion in the larger subparticles is of the same magnitude as that in the microparticles. These transfer resistances can, in some polymerization conditions, decrease polymer yield. Unfortunately, no predictions of MWD or CCD were reported. Some experimental results on particle morphology seem to support this model (Bukatov et al., 1982; Skomorokhov et al., 1987, Kakugo et al., 1989). Some mathematical models combine the approaches of the polymeric flow model and the multigrain model. In the polymeric multigrain model (Sarkar and Gupta, 1991), catalyst fragments are assumed to be in a continuum of polymer with only one level of diffusional resistance for the monomer. This model can predict higher values of polydispersity than the multigrain model. In the polymeric multilayer model (Soares and Hamielec, 1995b), the polymeric particle is divided into concentric spherical layers as in the multigrain model but microparticles are not considered. Therefore, the intraparticle radial profiles of monomer concentration and temperature for the multilayer model are the same as those of the polymeric flow model; i.e., mass- and heat-transfer processes that take place in the primary particles are not explicitly considered in the model. Prior to polymerization all layers have the same concentration of active sites as the whole catalyst particle; i.e., there is no radial profile of active sites. This assumption is supported by some recent electron microscopy studies of heterogeneous Ziegler-Natta catalysts (Noristi et al., 1994). Population balances for each active species are calculated inside each layer. The volume of each concentric layer is updated using average monomer concentrations and temperature after a given integration interval. The population balances derived for each layer are the same as those for a model with no transfer resistances. If transfer resistances are found to be of little importance, the same equations can still be used with the bulk monomer concentration and polymerization temperature. Additionally, the polymeric multilayer model estimates the distributions of molecular weight and chemical composition in each model layer and whole polymer using a theoretically sound equation, Stockmayers bivariate distribution (Stockmayer, 1945), as shown in Figure 7. The former expansion models could only estimate averages of molecular weight and composition. The instantaneous chain length and composition distribution proposed by Stockmayer and corrected for differing monomer molecular weights is given by

h h rv(polymeric flow) ) kpC*M

(209)

w(r,y) ) (1 + y)2r exp(-r) dr

Equations for the moments of the molecular weight distribution and copolymer composition have then to be

2/r exp(-y2r/2) dy (210)

992 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

Figure 7. Prediction of the bivariate distribution of chemical composition and chain length with the polymeric multilayer model. r ) chain length; W(r) dr ) chain-length distribution; y ) deviation from copolymer average composition; W(y) dy ) composition distribution; s1, s2, s3 site types; l1, l2, ..., l9 ) model layers; l1 is the innermost layer; l9 is the outermost layer.

Figure 8. Instantaneous chain-length distribution (CLD) of a polyolefin made with a multiple-site-type catalyst as a superposition of four individual Florys most probable weight chain-length distributions (solid line indicates CLD of whole polymer, and dotted lines represent CLD of polymer made on distinct active site types).

where

) F1(1 - F1)K h h h K ) [1 + 4F1(1 - F1)(r1r2 - 1)]0.5 h ) 1 - M2/M1 M2/M1 + F1(1 - M2/M1) h

(211) (212) (213)

and F1 ) average mole fraction of monomer type 1 on h copolymer, M1, M2 ) molecular weights of monomer types 1 and 2, respectively, r ) chain length, r1, r2 ) reactivity ratios, and y ) deviation from average copolymer composition. For single-site catalysts, the principal conclusions of the polymeric multilayer model confirm and complement the conclusions obtained with the polymeric flow model and the multigrain model: (1) Mass-transfer resistances can reduce the polymerization rate, decrease molecular weight averages, and affect copolymer composition for large, highly active catalysts. (2) Particularly important for supported catalysts, the concentration of highly active sites can increase mass-transfer resistances and have undesirable results in catalyst performance and product quality. (3) Mass-transfer resistances may also be a source of copolymer composition heterogeneity for highly active and large catalyst particles, if the comonomers have reactivities that differ significantly. (4) Temperature gradients in the polymeric particle are not expected to be a significant factor for reactions carried out in slurry reactors. The conclusions obtained with single-site-type expansion models are especially important for the technology of supported metallocene catalysts, where single-site, highly active catalytic species may be subjected to significant mass- and heat-transfer resistances (Soares and Hamielec, 1995a,b). Chemical Models. It is generally accepted that, under most polymerization conditions, the effect of multiple-site types on polymer properties is far more important than mass- and heat-transfer resistances (Ray, 1988; Soares and Hamielec, 1995b). Under these conditions, each site type instantaneously produces

Figure 9. Instantaneous chemical composition distribution (CCD) of a binary olefin copolymer made with a multiple-site-type catalyst as a superposition of five individual Stockmayers CCDs considering all chain lengths (solid line indicates CCD of whole polymer, and dotted lines represent CCDs of polymer made on distinct active sites).

polymer that has Florys most probable MWD. Therefore, the instantaneous MWD of accumulated polymer made with heterogeneous Ziegler-Natta catalysts can be considered an average of that produced by the individual site types, weighted by the weight fraction of polymer produced by each site type, as shown in Figure 8 (Soares and Hamielec, 1995c). The same treatment can be extended to copolymers by using Stockmayers bivariate distribution to describe the instantaneous bivariate distribution of chain length and composition. Therefore, if one assumes that each site type produces copolymer that obeys distinct bivariate distributions of chain length and copolymer composition, the bivariate distribution of the accumulated copolymer can be considered an average of that produced by the individual site types, as shown in Figure 9 (Soares and Hamielec, 1995e). If these hypotheses are valid, broad MWDs and CCDs result from the superposition of narrower distributions produced on each site type. Additionally, from the knowledge of the global MWDs and CCDs, one can use

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 993

Figure 10. CCDs for each site type over all model layers for propylene-ethylene copolymerization (Soares and Hamielec, 1995b).

Figure 11. CCDs considering all site types in different polymer layers for propylene-ethylene copolymerization. Layer 1 is the innermost layer, and layer 9 is the outermost layer (Soares and Hamielec, 1995b).

mathematical models to infer the number of active site types present on the catalyst (Vickroy et al., 1993; Soares and Hamielec, 1995c,e). Multiple-site-type models without heat- and mass-transfer resistances have achieved considerable success in simulating polymerizations using heterogeneous catalysts in slurry reactors (Rincon-Rubio et al., 1990; Lee et al., 1994), series of slurry reactors (de Carvalho et al., 1989, 1990; Soares, 1994), gas-phase reactors (McAuley et al., 1990; Xie et al., 1994), and high-pressure-high-temperature tubular reactors (Lorenzini et al., 1991). Hybrid Models. Hybrid models provide the most complete description of polymerization with heterogeneous Ziegler-Natta catalysts since they account for multiple-site types and transfer resistances simultaneously. Although multiple-site types are probably the main explanation for the characteristic behavior of heterogeneous Ziegler-Natta catalysts, transfer resistances may have a secondary effect on these properties in certain polymerization conditions (Ray, 1988; Soares and Hamielec, 1995b). Several of the expansion models described above have been modified to include the presence of more than one type of active site: the polymeric flow model (Galvan and Tirrell, 1986b), the multigrain model (Ray, 1988), the polymeric multilayer model (Soares and Hamielec, 1995b), and the polymeric multigrain model (Sau and Gupta, 1993). All these models indicate that the presence of more than one type of active site is mainly responsible for the broad MWDs and CCDs observed in polymer produced with heterogeneous Ziegler-Natta catalysts. In general, mass- and heat-transfer resistances are regarded as secondary effects. However, under certain polymerization conditions with large, highly active catalyst particles, mass-transfer resistances can have an important role in further broadening the MWD and CCD of polymers made with heterogeneous Ziegler-Natta catalysts. The polymeric multilayer model uses Stockmayers bivariate distribution to predict MWDs and CCDs of polymer chains per site type in each concentric layer of the model. These distributions are averaged over all particles to obtain the MWD and CCD of the whole polymer. Figure 10 shows the CCDs per site type of an ethylene-propylene copolymer made with a three-sitetype catalyst averaged over all model layers. Figure 11 shows the CCDs of the same copolymer per layer averaged over all site types. Finally, Figure 12 shows

Figure 12. Chain-length distribution for each site type over all layers for propylene-ethylene copolymerization (Soares and Hamielec, 1995b).

the MWDs per site type for the same polymer averaged over all model layers (Soares and Hamielec, 1995b). Among the expansion models, only the polymeric multilayer model can predict complete distributions of molecular weight and copolymer chemical composition. 3.1.2. Homogeneous Catalysts. Soluble ZieglerNatta catalysts, including metallocene catalysts, can make polymer with narrow MWD and CCD (Kaminsky, 1986a, 1991). This behavior supports the multiple-sitetype hypothesis for heterogeneous catalysts since these catalysts consist of reasonably well-defined, single catalytic species and should produce polymer that has narrow MWD and CCD. Some metallocenes, however, can synthesize polymer with broad MWD and CCD. This behavior has been linked to the presence of different site types (Kaminsky, 1986b). In fact, a combination of different types of metallocene catalysts can be used to produce polymer with broad MWD and CCD (Welborn, 1987, 1991, 1992, 1993), which is in good agreement with the chemical models presented above. In most processes using homogeneous Ziegler-Natta catalysts, in general, and metallocenes, in particular, the polymer is not soluble in the reaction medium and precipitates after a critical chain length is achieved. If after chain termination the active site returns to the

994 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

solution, then interparticle mass and heat transfer should not influence the polymerization. In this case, chemical models would be adequate to describe the polymerization. However, if the active sites are trapped inside polymer particles, interparticle mass- and heattransfer resistances could become significant. A mathematical model for particle growth during polymerizations catalyzed with metallocenes has been proposed recently (Herrmann and Bohm, 1991). It is possible that, as the polymer particles grow by agglomeration, some active sites will remain inside the polymer particles. This is especially more likely in processes where the active sites are exposed to high concentrations of monomer, as in liquid monomer polymerizations. In this case, due to the high activity of metallocene catalysts, transfer resistances may play a role in broadening MWD and CCD. Very few mathematical models have been proposed to describe polymerization using homogeneous ZieglerNatta catalysts. Soares and Hamielec (1995b) proposed that the polymeric multilayer model could be used to simulate polymerization with homogeneous ZieglerNatta catalysts if one neglects mass- and heat-transfer resistances in the polymeric particle. This would be rigorously valid only if active sites are not trapped inside polymer particles during polymerization. Vela-Estrada and Hamielec (1994) were the first to model the polymerization of ethylene with bis(cyclopentadienyl)zirconium dichloride and methylaluminoxane. Experimental polymerization rates and MWD results could be well represented by a two-site-type model. This was the first study to use size-exclusion chromatography to establish the number of active site types quantitatively for polymers synthesized with homogeneous Ziegler-Natta catalysts. In fact, a bimodal distribution gave a polydispersity index of 2 for each mode. This result is in excellent agreement with the model proposed in Figure 8. Supported metallocenes have also been studied extensively. Among the several incentives for supporting metallocenes are the use of already existing polymerization reactors for heterogeneous catalysts, the enhancement of stereochemical and regiochemical control, decrease in the methylaluminoxane/catalyst ratio, and production of polymer with better particle properties. The models described before for heterogeneous ZieglerNatta catalysts can also be applied to this class of catalysts (Soares and Hamielec, 1995b). In this case, in the presence of only one active site type on the catalyst, transfer resistances and statistical broadening would be the only factors causing broadening of MWD and CCD. Hoel et al. (1994) developed a simplified flow model, with no MWD prediction capabilities, and used it to interpret their experimental data on ethylene-propylene polymerization using two silica-supported metallocene catalysts in a liquid-propylene slurry reactor. Both catalysts produced polymer with narrow MWD at a constant rate of polymerization. However, one catalyst produced polymer with narrow CCD and the other produced polymer with broad CCD and decreasing amount of ethylene in the copolymer as a function of polymerization time (unfortunately, the catalyst types were not specified). The radial copolymer composition was measured by FTIR-microscope analysis for microtomed 250-750 m particles, and it was found to be richer in propylene in the center than in the exterior layers of the particle, which is also in good agreement

Figure 13. CCDs in different polymer layers for propyleneethylene copolymerization made on a single-site-type catalyst. Layer 1 is the innermost layer, and layer 10 is the outermost layer (Soares and Hamielec, 1995b).

with the multilayer model predictions for single-sitetype catalyst (Soares and Hamielec, 1995b), as shown in Figure 13. Although their model could qualitatively predict the radial profile of chemical composition, it could not account for the decrease in ethylene content during the polymerization. In reality, their model predicted an increase in ethylene content with polymerization time, and so would any other physical model published in the literature, since the concentration of active sites in the polymeric particle decreases during the polymerization, therefore decreasing the effect of intraparticle mass-transfer resistances. The observed decrease in ethylene content in the copolymer was tentatively attributed to variations in monomer diffusivity and particle porosity during the polymerization. Recently, Bonini et al. (1995) applied a modified form of the multigrain model to the polymerization of propylene with silica-supported Me2Si(Ind)2ZrCl2/MAO. Their model takes into account the mechanism of catalyst fragmentation by the growing polymer chains to explain the observed induction period when the supported metallocene is used. In their model, the catalyst particle is fragmented by the growing polymer in successive concentric shells, from the external surface to the center of the particle, as proposed by Ferrero and Chiovetta (1991). Their model agreed well with experimental rates of polymerization, including the induction period and number-average molecular weight, but could not predict the relatively broad MWD (polydispersities from 2.4 to 3.8). The simulated polydispersities were always less than 2.1, surprisingly low values. 3.1.3. Modeling of Long-Chain Branching. Constrained geometry metallocene catalysts can be used to synthesize polyolefins with narrow molecular weight distributions and significant degrees of long-chain branching. These polymers have excellent mechanical properties combined with good processability (good shear thinning, delayed melt fracture, and improved melt strength) (Hamielec and Soares, 1996). The most suitable catalyst types for long-chain branch formation appear to be those with an open metal active center, such as the Dow Chemical constrained geometry catalysts. The active center of these catalysts is based on group IV transition metals that are covalently bonded to a monocyclopentadienyl ring and bridged with

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 995

a heteroatom, forming a constrained cyclic structure with the transition-metal center. Strong Lewis acids are used to activate the catalyst to a highly effective cationic form. This geometry allows the titanium center to be more open to the addition of ethylene and higher R-olefins but also for the addition of vinyl-terminated polymer molecules (Lai et al., 1993a,b). The most likely long-chain branch formation mechanism with metallocene catalyst systems is terminal branching, a mechanism which has been known in the free-radical polymerization literature for many years. In free-radical polymerization, macromonomers are generated via termination by disproportionation and via chain transfer to monomer. With metallocene catalyst systems, the facile -hydride elimination reaction appears to be responsible for in-situ macromonomer formation. Other reaction types, such as -methyl elimination and trans, may also generate dead polymer chains with terminal unsaturation (Resconi et al., 1992). -Methyl elimination can actually be the most important transfer mechanism in propylene polymerization when (Me5Cp)2Ti, (Me5Cp)2Zr, and (Me5Cp)2Hf complexes are used (Huang and Rempel, 1995). Therefore, these catalytic systems have the potential of producing polypropylene with long-chain branches. The chain-length distribution of polymer chains produced with metallocene catalysts that permit long-chain branch formation via the terminal double-bond mechanism can be described analytically for each population containing a different number of long-chain branches per polymer molecule (Soares and Hamielec, 1996c). The polymerization kinetic model involves the following steps:

terminal vinyl unsaturation is proportional to the rates of -hydride elimination (producing dead polymer chains with terminal vinyl unsaturation) and transfer to chaintransfer agent (commonly to hydrogen, producing dead polymer chains with saturated chain ends). The frequency distribution of chain length for polymer populations with n long-chain branches per chain is given by (Soares and Hamielec, 1996c):

f(r,n) )

1 2n 2n+1 r exp(-r) (2n)!

(218)

where r represents chain length, n represents the number of long-chain branches per polymer molecule, and is given by

R RCTA RLCB + + Rp Rp Rp

(219)

where R is the rate of -hydride elimination, Rp is the rate of monomer propagation, RCTA is the rate of transfer to chain-transfer agent, and RLCB is the rate of macromonomer propagation or long-chain branch formation. The chain-length averages of the polymer populations with different numbers of long-chain branches per chain are related by the simple relationships:

j j rn,i ) (1 + 2i)rn,0 j rw,i ) (1 + i)rw,0 j j j rz,i ) (1 + 2i/3)rz,0 pdii ) +i (11+ 2i)pdi


0

(220) (221) (222) (223)

Pr,i + M 9 Pr+1,i 8
d 8 Pr,i + Dq,j 9 Pr+q,i+j kCTA kpLCB

kp

(214) (215) (216) (217)

8 Pr,i + CTA 9 Dr,i + P0,0 8 r,i Pr,i 9 Dd + P0,0


k

where Pr,i is a living polymer molecule of chain length d r containing i long-chain branches, Dq,j is a dead polymer molecule of chain length q containing j longchain branches and having terminal vinyl unsaturation, Dq,j is a dead polymer molecule of chain length q containing j long-chain branches and a saturated chain end, M is the monomer, CTA is a chain-transfer agent, kp is the propagation rate constant for monomer, kpLCB is the propagation rate constant for dead polymer with terminal vinyl unsaturation, kCTA is the rate constant for transfer to chain-transfer agent, and k is the rate constant for -hydride elimination. Chain initiation is assumed to be instantaneous. Dead polymer chains having terminal vinyl unsatd uration, Dq,j, can coordinate to the catalytic active site and insert in the growing chain, forming a trifunctional long-chain branch. Dead polymer chains with saturated chain ends, Dq,j, cannot polymerize again. Observe that by examining the mechanism of chain formation, one can conclude that the instantaneous molecular weight distributions of live polymer, dead polymer with vinyl chain-end unsaturation, and dead polymer with saturated chain ends will be the same. However, the relative amount of dead polymer with and without

where i indicates the number of long-chain branches per chain, and rn,i, rw,i, rz,i, and pdii are the number, weight, j j j and z-average chain lengths and polydispersity, respectively. Parts a and b of Figure 14 show the predicted weight chain-length distributions for a polyolefin produced in a CSTR for a given value of . Figure 15 shows the effect of varying the rate of long-chain branch incorporation in the chain-length distribution. The chainlength distribution becomes broader as the long-chain branching level increases, but the distribution is still very narrow as compared with the ones for polyolefins made with heterogeneous Ziegler-Natta catalysts. The analytical solution for the chain-length distribution of homopolymer with long-chain branches (eq 218) is also valid for copolymerization when appropriate pseudo-kinetic rate constants are used to calculate . Additionally, for the case of binary copolymerization, Stockmayers distribution can be used to obtain the chemical composition distribution of the copolymer. This is possible because the branched chains are formed by linear copolymer chains which follow Stockmayers bivariate distribution (Soares and Hamielec, 1996c). 4. Determination of Binary Reactivity Ratios The determination of reactivity ratios (terminal model) with small confidence intervals requires sensitive analytical techniques, carefully planned experiments, and the use of statistically valid methods of estimation. Unfortunately, most of the binary reactivity ratio data published to date were not found using statistically valid

996 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

Figure 14. Weight chain-length distribution for polymer populations containing different numbers of long-chain branches per polymer chain: (a) weight chain-length distribution normalized with respect to the weight of individual populations; (b) weight chain-length distribution normalized with respect to the total weight of the whole polymer. The global distribution (overall polymer populations) is indicated by the bold line. The table in the top right corner indicates number and weight fractions of each population.

methods and confidence intervals were usually not given. The reactivity ratios are highly correlated with major disagreement in values measured in different laboratories (ODriscoll and Reilly, 1987; Dube et al., 1991a; Burke et al., 1993; Rossignoli and Duever, 1995). 4.1. Statistically Valid Methods for Estimation of Reactivity Ratios. Traditional methods for estimating reactivity ratios (Mayo and Lewis, 1944; Fineman and Ross, 1950; Braun et al., 1973; Kelen and Tudos, 1975) are based on first transforming the in

stantaneous copolymer composition equation into a linear form in the parameters r1 and r2, for example

(f1/f2)(F2/F1 - 1) ) -r2 + r1(f1/f2)2(F2/F1) (224)


and then estimating the reactivity ratios by graphical plotting or by linear least squares. However, these approaches, aside from requiring that the instantaneous copolymer composition equation be valid, are statistically unsound because the independent variable

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 997

Figure 15. Effect of varying macromonomer addition rate on weight chain-length distribution.

(f1/f2)2(F2/F1) has error, and the dependent variable (f1/f2)(F2/F1 - 1) does not have constant variance. Both of the latter assumptions are necessary for linear least squares to be a statistically valid estimation method. As a result, they have been shown often to lead to very poor estimates with misleading confidence intervals (Tidwell and Mortimer, 1965, 1970). The common use of these statistically invalid estimation procedures in the past is one of the reasons for the wide variation in reactivity ratios reported by different researchers. It should be mentioned here that the Kelen and Tudos (1975) method, although statistically invalid, can be used to at least obtain good initial r1 and r2 estimates, provided the experiments have been suitably designed. More details are given in McFarlane et al. (1980) and Laurier et al. (1985). Reactivity ratios are nowadays most usually estimated using procedures based on the statistically valid error-in-variables model (EVM) or on its modifications (Box, 1970; Britt and Luecke, 1973; Sutton and MacGregor, 1977; Garcia-Rubio et al., 1985; Yamada et al., 1978; Van der Meer et al., 1978; Patino-Leal et al., 1980; Patino-Leal and Reilly, 1981; Dube et al., 1991a; Burke et al., 1993; Rossignoli and Duever, 1995). Extension of these EVM methods to the estimation of reactivity ratios in terpolymerizations has been discussed in Duever et al. (1983). These methods allow one to take properly into account all the sources of experimental error. No groups of variables are considered to be independent and free of error or dependent with constant error. All measured variables (e.g., [Mj]m, j ) 1,2 comonomers) are considered on an equal basis and represented as coming from some true (but unknown) value [Mj]t which is contaminated with measurement error, that is

The estimates obtained by these EVM approaches are superior to those obtained by arbitrary graphical or arbitrary least-squares procedures. Although EVM procedures require more information than arbitrary least-squares procedures, namely, the measurement error variances and covariances, the point estimates of the reactivity ratios obtained by these procedures have been shown to be very insensitive to changes in these covariance values. The great improvement provided by the EVM estimation methods comes from simply accounting correctly for the major sources of error. 4.2. Design of Experiments. For any experimental investigation, an optimal experimental design is of great importance since it enables one to perform the minimum number of experiments and obtain the most precise parameter estimates. In copolymerizations, such a design is that using the Tidwell-Mortimer (1965) criterion. The criterion employs the Mayo-Lewis model (e.g., see ODriscoll and Reilly (1987) or Hamielec et al. (1989)) and is based on the sensitivity of the reactivity ratio estimates to the errors encountered in the determination of copolymer composition. Tidwell and Mortimer (1965) recommended running several replicates at two different monomer feed compositions, f 1o and f 1o, defined by the following equations:

f ) 1o f 1o )

2 2 + r1 r2 2 + r2

(226)

(227)

[Mj]m ) [Mj]t + i i

(225)

It is evident from eqs 226 and 227 that preliminary estimates of the two reactivity ratios, r1 and r2, are needed. These estimates can be obtained from the literature (e.g., Polymer Handbook, 1989), a set of

998 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

preliminary experiments, a calculated guess using the Q-E scheme, etc. (ODriscoll and Reilly, 1987). In addition, the copolymerizations should be carried out to a conversion low enough that composition drift (i.e., a change in the monomer feed composition, f10) is negligible (e.g., conversion <5-8%). 5. Alternative Copolymerization Kinetic Models Deviations from the terminal model have been noted for various comonomer pairs and for various polymerization systems including free-radical, anionic, and heterogeneous Ziegler-Natta. In order to explain some deviations from the terminal model in free-radical polymerization, several alternative models have been proposed to describe copolymerization kinetics (Harwood, 1987). A model first formulated by Merz et al. (1946), the penultimate unit effect model has received much attention over the last decade. This model describes how the reactivity of a radical center may depend on the type of monomer unit bound in the chain end adjacent to the radical center. Much of the recent interest in the penultimate model is due to the work of Fukuda et al. (1985, 1987, 1989, 1991). Using the pseudo-kinetic rate constant modeling approach described earlier, extension of these alternative models to multicomponent systems is straightforward. With the existence of several alternatives to describe the polymerization mechanism, it is then left to distinguish between the models to find which gives the best description of experimental data. Early model discrimination attempts were geared toward fitting models to measurements of copolymer composition and diad and triad fractions (Hill et al., 1982, 1984, 1989; Brown et al., 1993). Some researchers have proposed that copolymer composition data can usually be adequately represented by terminal model kinetics since differences between model predictions are typically small (Fukuda et al., 1987). On the other hand, results by Burke et al. (1994a) have indicated that, with the use of statistical experimental design techniques, these data may be more useful in discriminating between copolymer models than previously thought. However, sequence distribution and copolymer composition data do not provide enough information to estimate all of the parameters contained in some models (e.g., the penultimate model). Thus, in recent attempts to discriminate between competing copolymerization models, polymerization rate data were used to gain additional information on model adequacy (Fukuda et al., 1985; Davis et al., 1989). Typical attempts at identifying the best model have involved spreading experiments over the entire range of feed compositions. This may not be optimal for either model discrimination or parameter estimation (Burke et al., 1994a). Rather, the use of statistical model discrimination methods which include the design of experiments and the measurement of copolymer composition, triad fractions, and copolymerization rate is preferred (Burke et al., 1994a,b, 1995, 1996). The results of Burke et al., which were verified experimentally, showed that the use of model discrimination techniques should improve our ability to discriminate between proposed copolymerization kinetic models. 6. Semibatch Monomer Feed Policies In the absence of an azeotrope and when one monomer is more reactive than the other in a binary, batch copolymerization (say r1 > 1 and r2 < 1), the instanta-

neous copolymer composition will decrease in monomer 1 with an increase in monomer conversion. The extent of compositional drift which leads to a copolymer heterogeneous in composition depends on the ratio of reactivity ratios r1/r2 (increase with an increase in r1/r2), the initial monomer composition (f10), and the monomer conversion (x). A copolymer which is heterogeneous in composition usually has inferior properties, and therefore industrial semibatch processes have been developed to reduce composition hetereogeneity. There are, however, certain semibatch emulsion polymerization processes where heterogeneous copolymers are produced in the latex particles to achieve certain property improvements (Bassett and Hamielec, 1981). There are two basic monomer feed policies, originally shown in Hamielec et al. (1987), which may be used in semibatch copolymerization to minimize compositional drift. Effective commercial processes are usually based on one or a combination of these feed policies. Some additional promising derivations of these policies have been presented (Arzamendi and Asua, 1989, 1990, 1991; Arzamendi et al., 1991, 1992; van Doremaele et al., 1992; Leiza et al., 1993; Schoonbrood et al., 1993, 1996; Canegallo et al., 1994; Canu et al., 1994; Urretabizkaia et al., 1994a; de la Cal et al., 1995; Echevarria et al., 1995). These feed policies (Policy I and Policy II) are described below using binary copolymerization as an example. Policy I. All of the slower monomer and sufficient enough of the faster monomer (to give the desired composition F1) are added to the reactor at time zero. Thereafter, the faster monomer is fed to the reactor with a time-varying feed rate to maintain N1/N2 (the ratio of the number of moles of monomer 1 to that of monomer 2 in the reactor) and F1 constant with time or grams of copolymer produced. Policy II. A heel of monomers 1 and 2 at the desired concentration levels (to give the desired F1) is added to the reactor at time zero. Thereafter, monomers 1 and 2 are fed to the reactor with time-varying feed rates to maintain [M1], [M2] and F1 constant with time or grams of copolymer produced. The equations to be solved to determine the required feed rates to produce a homogeneous binary copolymer in a semibatch process are given by eqs 228-233.

dN1 ) -N1(k111 + k212)[P] + F1,in dt dN2 ) -N2(k121 + k222)[P] + F2,in dt

(228) (229)

1 1 dV F1,inMW1 F2,inMW2 ) + - Rp1MW1 + dt Fm1 Fm2 Fm1 Fp 1 1 V (230) Rp2MW2 Fm2 Fp

)]

Initial Conditions t ) 0, N1 ) N1o, N2 ) N2o, V ) Vo Conditions Policy I: F2,in ) 0 and d(N1/N2) )0 dt (232) (231)

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 999

Policy II: d[M1] d[M2] ) )0 dt dt (233)

and

2F1k**kp[M]0Yo p kfCTA[CTA]

(243)

Since f1 and F1 are constant, 1 is also constant. Given the time variation of the total polymer radical concentration, [P], one can readily solve for F1,in and F2,in, the time-varying monomer feed rates, using the above equations. We will now investigate Policies I and II with respect to cross-linking using the criterion of onset of gelation already discussed. Monomer 1 is the cross-linker (divinyl monomer) and a chain-transfer agent (CTA) is used to delay the onset of gelation. 6.1. Policy I. For constant [P] and F2,in ) 0, eq 229 may be solved analytically for N2 to give

Q2 goes to infinity at the gelation point t ) tc and tc is given approximately by

tc ) 1/ )

kfCTA[CTA] 2F1k**kp[M]0Yo p

(244)

N2 ) N2oe-2t
where

(234)

2 ) (k121 + k222)[P]
For constant N2/N1, N1 is given by

(235)

N1 ) N1oe-2t
One may now solve eq 228 for F1,in to get

(236)

We employ the following set of kinetic parameters: kfCTA ) 102 L mol-1 s-1, [CTA] ) 0.1 mol L-1, a constant k** p ) 102 L mol-1 s-1, F1 ) 0.01, kp ) 103 L mol-1 s-1, Y0 ) 10-7 mol L-1, [M]0 ) 10 mol L-1, and tc ) 5000 s (1.4 h). Employing eqs 159, 160, and 176 for this example of Policy I, it can readily be shown that, at the gelation point (t ) tc), BN4 = 1/3. In other words, at the gelation h point one polymer molecule in three on the average has a tetrafunctional long-chain branch point. 6.2. Policy II. To simplify the analysis, we assume that the densities of monomers and copolymer are the same and that the molecular weight of the monomers are the same, in addition to constant [P]. Equation 229 may be solved analytically to give the following:

N1 ) N1oeRt (237) N2 ) N2oeRt F1,in ) (R + 1)N1oeRt F2,in ) (R + 2)N2oeRt


where

(245) (246) (247) (248)

F1,in ) N1o(1 - 2)e-2t

Assuming that the densities of monomers and copolymer are the same (Fm1 ) Fm2 ) Fp), one can integrate eq 230 analytically to get

N1oMW1 1 - 2 V ) Vo + (1 - e-2t) (238) Fm1 2


The total monomer concentration in the reactor is therefore given by

))

R)

f1o1 + f2o2 (MW(N1o + N2o) - 1) VoFm

(249)

[M] )
where

N1 + N2 (N1o + N2o)e-2t ) V V + R(1 - e-2t)


o

(239)

Since the ratio F1,in/F2,in is independent of time, the two monomers may be premixed and fed to the reactor in the same stream.

V ) VoeRt R ) N1oMW1 1 - 2 2/Fm1 (240)

(250)

We will now investigate gelation for a binary copolymerization of vinyl and divinyl monomers. The divinyl monomer is considered the more reactive monomer, and therefore, with Policy I, it is fed into the reactor over time. Low levels of the cross-linking monomer are to be used and therefore to simplify the analysis, one can assume that the volume of the reacting mixture, V, is constant and equal to the initial volume, V0. We now apply eq 161 for constant V, ) Cp1 ) Cp2 ) 0 and ) kfCTA[CTA]/(kp[M]) with Policy I to give

We will now investigate gelation for a binary copolymerization of a vinyl and divinyl monomer. To simplify the analysis, it is assumed that the concentration of chain-transfer agent, [CTA], is maintained constant. We now apply eq 161 under the following conditions: ) Cp1 ) Cp2 ) 0 and ) kfCTA[CTA]/ (kp[M]) constant with time in the context of Policy II.

dQ2/dt ) (1 + Q2)2 - RQ2 h h

(251)

h dQ2/dt ) (1 + Q2)2e-2t h
with Q2 at t ) 0 where h

with Q2 ) 0 and t ) 0. The definitions of Q2 and are h h the same as those used with eq 241. Equation 251 can be solved numerically (setting ) 4.54 10-4 s-1 and R ) 1.515 10-4 s-1) to give
t (s) 0 Q2 0 h 100 200 300 400 500 700 900 0.0472 0.0983 0.154 0.215 0.282 0.438 0.636 1500 1.769 2310 91.1 1700 2.61 1900 4.22 2100 8.57 2150 2200 11.13 15.60

(241)

Q2 ) h

k**Q2 p kp[M]

t (s) 1100 1300 Q2 0.894 1.249 h

(242)

t (s) 2250 Q2 25.4 h

2300 64.0

2320 2330 157.0 556.0

1000 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

Neglecting RQ2 in eq 251, one gets an analytical solution h for Q2 h

Q2 ) h
and

t 1 - t

(252)

tc ) 1/

(253)

Using eq 253, one finds tc ) 2203 s. RQ2 in eq 251 h accounts for the volume increase due to monomer fed and has the effect of delaying the onset of gelation. The following set of kinetic parameters has been used to calculate and R and will be used to calculate the number- and weight-average molecular weights and branching frequency (BN4) with Policy II. h

r1 ) 2, r2 ) 0.5 (monomer 1 is the divinyl monomer), F1 ) 0.2 k11 ) k22 ) 2 103 L mol-1 s-1 1 ) 0.333, f1 ) 0.1111, 1 ) 3.33 10-4, 2 ) 1.667 10-4 s-1 Fm ) Fm1 ) Fm2 ) Fp ) 1000 g L-1 MW ) MW1 ) MW2 ) 100 Vo ) 1 L, N1o ) 0.5 mol, N2o ) 4 mol [M1] ) 0.5 mol L-1, [M2] ) 4 mol L-1 kp ) 2.52 103 L mol-1 s-1, k** ) 10 L mol-1 s-1 p kfCTA ) 102 L mol-1 s-1, [CTA] ) 0.1 mol L-1, Yo ) 10-7 mol L-1 R ) 1.515 10-4 s-1, ) 4.54 10-4 s-1
In the context of Policy II, one can solve eqs 159, 160, and 176 analytically to give

infinite; however, one can continue to calculate rN and j BN4 beyond the point of gelation. Experimental data h clearly show that once gel is formed, it grows very rapidly at the expense of the sol. The gel acts like a great sponge, sucking in the polymer from the sol. Radical centers located on polymer chains in the gel are longer lived because of their greatly reduced mobility. These radical centers are terminated by reaction diffusion (these radical centers move in the gel by the act of monomer addition or propagation). The greatly reduced termination rates of radical centers located on chains in the gel result in a very high concentration of these radical centers, and this in part is the cause of the rapid growth of the gel at the expense of the sol. One can identify the following reactions as responsible for the growth of the gel. 1. k**Y1,b[Rs] consumption rate of polymer radicals p in the sol by addition to double bonds bound in polymer chains in the gel (Y1,b equals the concentration of monomer bound in the gel and [Rs] is the concentration at radical centers on the gel). 2. kp[M][Rb] consumption rate of monomer by radical centers located on polymer chains in the gel. 3. k**Q1,s[Rb] addition rate of polymer chains from p the sol to radical centers located on the chains in the gel (Q1,s is the concentration of monomer bound in the polymer chains in the sol). 4. ktsb[Rs][Rb] termination rate of polymer radicals in the sol with those on the gel. Termination of radical centers on the gel may result from the following reactions in addition to number 4 above: ktrd[Rb]2 termination rate of radical centers on chains in the gel by reaction diffusion (this reaction type produces intramolecular cross-links and leads to a denser gel with lower absorbency and higher gel strength). As soon as gel is formed, it grows rapidly at the expense of the sol with the above reactions playing a major role in its rapid growth. 7. Some Recent Case Studies 7.1. Introduction. In the previous review by Hamielec et al. (1987), an evaluation of the free-volume theory for the modeling of diffusion-controlled termination and propagation in binary copolymerization as well as the pseudo-kinetic rate constant method was presented. Several case studies involving the implementation of these ideas are summarized in Table 1. Early application of the pseudo-kinetic rate constant method was made for the batch, semibatch, and continuous emulsion polymerization of styrene/butadiene rubber by Broadhead et al. (1985). Four other comonomer systems (styrene/acrylonitrile, styrene/p-methylstyrene, p-methylstyrene/acrylonitrile, and p-methylstyrene/methyl methacrylate), modeled using the free-volume theory to account for diffusion-controlled termination and propagation in batch reactors (Garcia-Rubio et al., 1985; Bhattacharya and Hamielec, 1986; Jones et al., 1986; Yaraskavitch et al., 1987), were discussed in detail in Hamielec et al. (1987). Since then, Storti et al. (1988, 1989) employed the pseudo-kinetic rate constant method to model the emulsion copolymerization of styrene/ acrylonitrile and acrylonitrile/methyl methacrylate as well as the emulsion terpolymerization of styrene/ acrylonitrile/methyl methacrylate. Xie and Hamielec (1993c) modeled solution polymerizations with added

VQ0 ) (F1k**[M]Yo2Vo/R)[t - (1/R)(eRt - 1)] + p (1/R)(kfCTA[CTA]YoVo)(eRt - 1) (254) VQ1 ) (kp[M]YoVo/R)(eRt - 1) BN4 ) h (F1k**[M]Yo2Vo/R)[(1/R)(eRt - 1) - t] p VQ0 (255) (256)

Employing the above kinetic parameters in eqs 254, 255, and 256 and eqs 155 and 156, one can calculate rN, and j j rW, and BN4 versus time of polymerization to give h
t (s) j rN j rW BN4 h 0 1134 2270 0 100 1147 2380 0.0113 500 1201 2930 0.0592 1000 1274 4070 0.1242 2000 1445 16 650 0.274 2500 1544 0.362

j j At this point of gelation, t ) tc ) 2350 s, rN ) 1513, rW ) , and BN4 ) 0.333. h 6.3. Beyond the Point of GelationsGel Growth and Structure. Beyond this critical point, the calculation of Q2 and rW is no longer possible as they are both j

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 1001


Table 1. Some Examples of Practical Applications of Copolymerization Models Broadhead et al., 1985 Garcia-Rubio et al., 1985 Bhattacharya and Hamielec, 1986 Jones et al., 1986 Yaraskavitch et al., 1987 Storti et al., 1988 Penlidis et al., 1989 Storti et al., 1989 Xie et al., 1991 Xie and Hamielec, 1993c Gao and Penlidis, 1996 batch, semibatch, and continuous emulsion styrene (Sty)/butadiene bulk Sty/acrylonitrile (AN) bulk Sty/p-methylstyrene bulk p-methylstyrene/MMA bulk p-methylstyrene/AN emulsion Sty/AN, AN/MMA, Sty/AN/MMA solution MMA emulsion AN/Sty/MMA suspension VC solution Sty/ethylene dimethacrylate, Sty/divinylbenzene bulk and solution MMA, Sty, AN, BA, ethyl acrylate, methyl acrylate, VAc, p-methylstyrene, carboxylic acid

Figure 16. Bulk Sty/EA: model predictions of conversion vs time data from McManus and Penlidis (1996) for [AIBN] 0.05 M at 50 C.

Figure 17. Bulk Sty/EA: model predictions of conversion vs time data from McManus and Penlidis (1996) for [AIBN] 0.05 M at 60 C.

cross-linker (styrene/ethylene dimethacrylate, styrene/ divinylbenzene) using similar techniques. Further applications of the free-volume theory were demonstrated by Xie et al. (1991) for the suspension polymerization of vinyl chloride and by Penlidis et al. (1992) for the solution polymerization of MMA. The latter showed how to simplify the general model (similar to the one described herein) for a particular case. Of most significant note, however, is the work by Gao and Penlidis (1996) which demonstrated the general applicability of the modeling techniques presented herein. They modeled bulk and solution polymerizations of several monomers as well as some suspension polymerization cases (see Table 1). The most convincing aspects of that work were the good model predictions of data from a large number of research groups and widely varying reaction conditions. 7.2. Styrene/Ethyl Acrylate Model Prediction Case Study. An experimental study of the bulk freeradical copolymerization of styrene (Sty)/ethyl acrylate (EA) was carried out by McManus and Penlidis (1996). This thorough kinetic study included conversion, composition, and molecular weight data for reactions at a variety of feed compositions, initiator concentrations, and temperatures, with and without added chaintransfer agent. The model described herein was applied to these data, and some representative results are shown below. In Figures 16 and 17 are shown model predictions which correctly predict the conversion vs time data for three different monomer feed compositions at 50 and

Figure 18. Bulk Sty/EA: model predictions of copolymer composition vs conversion data from McManus and Penlidis (1996) for [AIBN] 0.05 M at 60 C.

60 C, respectively, for a given initiator concentration. The cumulative copolymer composition is plotted against conversion in Figure 18 for three different monomer feed compositions at 60 C. These results are representative of all of the composition plots. The good agreement between the composition data and the model predictions

1002 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

Figure 19. Bulk Sty/EA: model predictions of number- and weight-average molecular weight vs conversion data from McManus and Penlidis (1996) for [AIBN] 0.05 M at 60 C.

Figure 21. Bulk BA: model predictions of conversion vs time data from Dube et al. (1991b) at 60 C.

Figure 20. Bulk Sty/EA: model predictions of number- and weight-average molecular weight vs conversion data from McManus and Penlidis (1996) for [AIBN] 0.10 M at 60 C.

is not surprising since the reactivity ratios used in this study were estimated using the statistically valid methods outlined earlier. Finally, representative plots of number- and weightaverage molecular weight vs conversion are shown in Figures 19 and 20 for two different initiator concentrations at 60 C. The model correctly predicts the trend of increased molecular weight as a result of an increase in the initiator concentration. The important point to consider in this case study is that only one set of database parameters was used to predict the data over a wide variety of reaction conditions. 7.3. BA/MMA/VAc Model Prediction Case Study. A comprehensive study of the butyl acrylate (BA)/methyl methacrylate (MMA)/vinyl acetate (VAc) terpolymer system was recently reported in Dube (1994) and Dube and Penlidis (1995a-c, 1997). While the eventual objective of the work was to develop an understanding of the emulsion terpolymerization system, bulk, solu-

tion, and emulsion homo- and copolymerization data were also collected. The model described herein was applied to these systems, and some representative results are discussed below. The parameter values used in the model are found in Dube, 1994. As an example of a bulk homopolymerization, the prediction of bulk BA data from Dube et al. (1991b), is shown in Figure 21. From these data, values for the propagation and termination rate constants (kp and kt, respectively) were estimated using only one set of data, hence, more accurate values of these rate constants are required (see Dube, 1994). Updated values of these rate constants are now available from pulsed-laser polymerization studies (Buback et al., 1989, 1994; Buback and Degener, 1993; Lyons et al., 1996). Values for the freevolume parameters A and VFpcrit were fit to the data at one set of reaction conditions. For the BA system, rate constants for transfer to polymer and terminal doublebond reactions are unknown, thus necessitating the acquisition of molecular weight data. This is particularly difficult due to the presence of significant amounts of gel in the samples (Dube et al., 1991b). Although no molecular weight plots are shown for these data, it was noted that Mn and Mw are extremely sensitive to the h h value of kfm. As mentioned earlier, a thorough investigation of bulk and solution modeling showing a large set of examples (including BA, MMA, and VAc bulk homopolymerization) is found in Gao and Penlidis (1996). Bulk copolymerization data reported in Dube and Penlidis (1995a) (the reaction conditions can be found therein) are plotted along with model predictions in Figures 22-25. Conversion vs time data for bulk BA/ MMA runs are plotted in Figure 22. The free-volume parameters A, K3, and VFpcrit were fit to data at one set of reaction conditions. The reader is referred to Dube and Penlidis, 1995a, for the model prediction of copolymer composition. Molecular weight data are shown in Figure 23. No parameter fitting to the molecular weight data was attempted. As was the case for the bulk homopolymerization modeling of BA, the molecular weight predictions were quite sensitive to the value for the transfer to monomer constant, kfm, for BA. Results from a bulk BA/VAc and a bulk MMA/VAc run are

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 1003

Figure 22. Bulk BA/MMA: model predictions of conversion vs time data from Dube and Penlidis (1995a) for f10 ) 0.439 at 60 C.

Figure 24. Bulk BA/VAc: model prediction of conversion vs time data from Dube and Penlidis (1995a) for f10 ) 0.8, [AIBN] ) 0.00054 mol/L at 60 C.

Figure 23. Bulk BA/MMA: model predictions of molecular weight vs conversion data from Dube and Penlidis (1995a) for f10 ) 0.439, [AIBN] ) 0.005 mol/L at 60 C.

Figure 25. Bulk MMA/VAc: model prediction of conversion vs time data from Dube and Penlidis (1995a) for f10 ) 0.3, [AIBN] ) 0.01 mol/L at 60 C.

shown with model predictions in Figures 24 and 25, respectively. As was the case for the BA/MMA runs, the free-volume parameters A, K3, and VFpcrit were fit to the data. The prediction of the composition data for both BA/VAc and MMA/VAc are reported in Dube and Penlidis, 1995a. Values for the free-volume parameters A, K3, and VFpcrit were fit to the BA/MMA/VAc conversion data from Dube and Penlidis, 1995b, for one set of reaction conditions. Conversion vs time data and model predictions for the bulk terpolymerizations are plotted in Figures 26 and 27. The prediction of the composition data is plotted in Figure 28. The predictions of the terpolymer composition data were achieved through the use of copolymerization reactivity ratios estimated using the optimal methods described earlier. Molecular weight predictions also followed similar trends (see Dube, 1994).

In the prediction of emulsion polymerizations, a larger number of physical constants and rate parameters were required. It was not possible to obtain all of the constants from the open literature, nor was it possible to obtain empirical estimates of the parameters. Therefore, many of the parameters used in the prediction of the emulsion polymerization data were arrived at by way of calculated guesses. In several cases, unknown parameters were given values previously reported for other monomers (see Dube, 1994). Prediction of conver sion for the BA emulsion polymerization run described in Dube and Penlidis (1995c) is plotted in Figure 29. Despite the uncertainty in the various rate parameters for BA, the model prediction was quite good. Emulsion terpolymerization composition vs conversion data are shown in Figures 30 and 31. This is yet another confirmation of the accuracy of the reactivity ratios as well as the partition coefficients used in the model (Dube, 1994).

1004 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

Figure 26. Bulk BA/MMA/VAc (30/30/40 wt %): model predictions of conversion vs time data from Dube and Penlidis (1995b) at 50 C.

Figure 28. Bulk BA/MMA/VAc (30/30/40 wt %): model predictions of terpolymer composition vs conversion data from Dube and Penlidis (1995b).

Figure 29. Emulsion BA: model prediction of conversion vs time data from Dube et al. (1995c). Figure 27. Bulk BA/MMA/VAc (30/30/40 wt %): model predictions of conversion vs time data from Dube and Penlidis (1995b) at 70 C.

VQ ) [(k111 + k212)N1(-H1) + (k121 + k222)N2(-H2)] (257)

7.4. Semibatch Production of Multicomponent Polymers. The practical implementation of monomer feed policies requires the use of on-line (or possibly offline) measurements to permit one to adjust for uncontrolled variations in recipe impurities such as O2 (which is a radical scavenger at temperatures below about 100 C) and other impurities in the recipe components which can affect radical concentration. An extensive review of on-line sensors for polymerization reactors was published by Chien and Penlidis (1990). Examples of on-line calorimetry, gas chromatography, and densitometry are discussed below. 7.4.1. Calorimetric Control of Monomer Feed in a Copolymerization. The instantaneous heat generation due to polymerization (VQ) is given by

where V is the volume of polymerizing mixture, Q is the instantaneous heat generation rate due to polymerization in cal L-1 s-1, -H1 is the heat of polymerization when monomer 1 adds to either polymer radical type and similarly for -H2. The feed rate of M1, F1,in, is given by

F1,in )

1 - 2

1 - R(N1/N2)

N1

(258)

for any monomer feed policy, where

R ) F2,in/F1,in

(259)

Dividing eq 257 by eq 258 and thus eliminating the total polymer radical concentration, one obtains

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 1005

Figure 30. Emulsion BA/MMA/VAc (30/30/40 wt %): model predictions of terpolymer composition vs conversion data from Dube (1994).

Figure 31. Emulsion BA/MMA/VAc (35/15/50 wt %): model predictions of terpolymer composition vs conversion data from Dube (1994).

VQ ) F1,in (k111 + k212)(-H1) + (k121 + k222)(-H2)(N2/N1) (k11 - k12)1 + (k21 - k22)2 1 - R(N1/N2) (260)

If the two monomers are premixed and fed to the reactor in one stream, R is a constant. It should be recalled that 1 and 2 are both functions of N1/N2 and polymerization temperature. 1 may be expressed as

1 )

1 1 + k12N2/k21N1

(261)

The temperature dependence of 1 should be small because the activation energies of the cross-propagation constants are similar in magnitude. In fact, the right-

hand side of eq 260 should have a small temperature dependence. In other words, to maintain constant F1 for the copolymer being produced, one should control the monomer feed rate (F1,in or (1 + R)F1,in) to maintain VQ/F1,in constant with time. The on-line measurement of VQ is obviously a convenient method for the control of copolymer composition with any monomer feed policy (N.B. when R ) 0, Policy I is being employed). On-line measurement of VQ can also be used to achieve a desired variation of N1/N2 and spread in copolymer composition. To maintain the desired production level of copolymer, one can increase polymerization rate by increasing the total polymer radical concentration, [P]. This can be done by increasing the reaction temperature while maintaining VQ/F1,in constant. Hendy (1975) used calorimetric control to produce high AN copolymers of styrene/acrylonitrile of uniform composition in emulsion polymerization. Two monomer feeding methods were employed, and both provided composition control. In one, a monomer mixture with the same composition as that desired for the copolymer was fed to the reactor (Policy II with monomer-starved feeding), and with the second feeding method, only the more reactive monomer was fed to the reactor (Policy I). For the high AN copolymers synthesized by Hendy, it was found that Policy I (feeding in styrene alone) gave better copolymer products than Policy II (with monomer-starved feeding). The specific reasons for this were not given. In the production of AN/S copolymers by feeding styrene, a constant feed rate of styrene was employed and this appeared to give uniform composition copolymer at least up to 80% conversion. Samples were withdrawn during the polymerization for analysis, and it was found that at about 80% conversion of AN the copolymer was slightly richer in styrene (2% higher in styrene) than the copolymer produced up to 50% conversion. This was attributed to the partitioning of AN into the water phase (10% soluble in water) and the effect on the AN/S ratio in the polymer particles at higher conversions. Equation 260 should only apply to stage III in emulsion polymerization when both monomers have very low solubilities in the water phase. Hoffman (1984) found similar results. Moritz (1989) reviewed the state of the art of isothermal bench-scale calorimeters and their application to polymerization reactions. He developed a microcomputer bench-scale polymerization calorimeter with online determination of polymerization rate and monomer conversion and studied the batch and semibatch emulsion polymerization of vinyl acetate. Lately, the use of on-line reaction calorimeters in emulsion polymerization has been gaining favor because on-line gas chromatograph setups allow only for discrete and delayed measurements. An automated reaction calorimeter was used to monitor the rate of emulsion polymerization of styrene using different emulsifier and initiator concentrations in conjunction with off-line measurements of the evolution of the particle-size distributions by de la Rosa et al. (1996). Their experimental results suggested that the end of nucleation and the disappearance of monomer droplets take place at approximately the same conversion (x = 36-42 wt %). A method to determine the minimum time monomer addition policy for composition control in the semibatch unseeded emulsion copolymerization of BA/VAc using calorimetric measurements was developed by Gugliotta et al. (1995b). An iterative

1006 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997

empirical approach using semistarved conditions was used. This was accomplished using pure and technicalgrade monomers. Saenz de Buruaga et al. (1996) presented further results for the same system. Case studies involving the emulsion copolymerizations of BA/ MMA, BA/VAc, and MMA/VAc were examined by Urretabizkaia et al. (1993) using a similar setup. 7.4.2. On-Line Monomer Concentration Measurements for Control of Monomer Feed. On-line measurements of the monomer ratio N1/N2 in the polymerizing phase (polymer particles in emulsion polymerization) with a feedback system to adjust monomer feed rates can be used to control copolymer composition. Guyot et al. (1981) used an on-line GC to control copolymer composition in the emulsion copolymerization of styrene/acrylonitrile. They employed Policy I but found in their subsequent analysis of the copolymers that copolymer composition drifted toward higher styrene levels. They showed that this was due to significant partitioning of acrylonitrile into the water phase. This was not corrected for in the monomer feed rates employed to control copolymer composition at a uniform level. Askill and Gilding (1981) produced uniform composition copolymers of methyl methacrylate/acrylic acid (rMMA 2.3 and rAA 0.31) using semibatch solution polymerization and Policy II (with monomer-starved feed). The solvent used was butanone. On-line GC measurements were used to monitor the monomer ratio (NMMA/NAA) in the reactor and adjust monomer feed rates. For most of the semibatch runs under monomerstarved conditions, there was no need to adjust feed rates. Narrower molecular weight distributions were found for the uniform composition copolymers produced by use of Policy II (with monomer-starved feed). This confirms that reactions which form long trifunctional branches such as transfer to polymer and reaction with terminal double bonds are insignificant for the MMA/ AA system. Guyot et al. (1984) monitored monomer concentrations by GC and used semibatch copolymerization of butadiene/acrylonitrile to obtain uniform composition polymers (Policy I). Constant composition copolymers showed a single glass transition temperature, and the production of cross-linked gel was delayed and sometimes avoided. The production of gel occurred due to polymer radicals adding to internal double bonds in l,2-butadiene units bound in the copolymer chains. Reactivity ratios were estimated to be rB ) 0.3 and rA ) 0.04. Semibatch copolymerization not only produced copolymers uniform in composition but also polymer with lower levels of tetrafunctional long-chain branching and cross-linked gel. The reactive internal double bonds have the vinyl 1,2-structure, and these were responsible for the formation of cross-linked gel. In butadiene homopolymerization about 20% of the internal double bonds have the 1,2-structure. Apparently, during semibatch synthesis of uniform copolymer the incorporated acrylonitrile orientated the added butadiene unit to the less reactive 1,4-structure because the sequence length of butadiene was kept small. In batch copolymerization, acrylonitrile was consumed preferentially, and at high conversions, copolymer containing long butadiene sequences and higher levels of vinyl 1,2-structure was formed. This, of course, was responsible for the formation of the high level of cross-linked gel. Snuparek and Kaspar (1981) investigated the semibatch emulsion copolymerization (Policy II with mono-

mer-starved feed) of ethyl acrylate/butyl acrylate. The copolymerization was followed by gas chromatography. There is evidence that the copolymerization may be controlled by monomer diffusion in the polymer particles, with both reactivity ratios approaching unity. El-Aasser et al. (1983) investigated the batch and semibatch emulsion copolymerization of vinyl acetate/ butyl acrylate. Policy II with monomer-starved feed was employed. Chain transfer to vinyl acetate bound in the copolymer chains produced long trifunctional branches and led to broader molecular weight distributions which were bimodal. Leiza et al. (1993) used an on-line GC in their development of a semiempirical approach to determine minimum time optimal monomer addition policies to produce homogeneous methyl methacrylate/ethyl acrylate copolymers. The method involves a series of semibatch emulsion copolymerizations carried out under semistarved conditions. The composition control of a semibatch BA/MMA/VAc emulsion terpolymerization was achieved with an online GC by Urretabizkaia et al. (1994b). Their strategy is based on comparing on-line measurements to model predictions and using the difference as input to a nonlinear adaptive plus proportional-integral controller. The controller calculates the feed rates of the more reactive monomers that have to be fed into the reactor to make sure that, after a sampling time interval, the monomer ratios in the polymer particles lead to the formation of a terpolymer of the desired composition. Experimental verification was performed using a 55 wt % solids content and a molar terpolymer composition target of BA/MMA/VAc ) 50/35/15. 7.4.3. On-Line Density Measurements for Composition Control. The use of on-line density measurements has gained increasing importance as attested by some earlier references in Chien and Penlidis (1990). More recently, Canegallo et al. (1993) used an on-line densitometer coupled with a kinetic model to monitor conversion and copolymer composition in batch emulsion homopolymerizations (Sty and MMA) as well as copolymerizations (Sty/MMA, AN/MMA, and MMA/VAc). Canu et al. (1994) then developed an a priori monomer feed policy which requires implementation in conjunction with the densitometer described above. Experimental verifications with emulsion copolymerizations (MMA/VAc, Sty/BA) and a terpolymerization (BA/MMA/ VAc) were reported. 7.4.4. Policies I and II in Semibatch Emulsion Copolymerization. It should be mentioned that the strict application of Policies I and II is not always possible in emulsion copolymerization. In stages I and II, monomer droplets are also feeding the polymer particles (the major site for polymerization), and this must be properly accounted for when calculating the required feed rates of the monomers, F1,in and F2,in. In stage III, however, if the solubilities of the monomers in the water phase are negligible, Policies I and II can be strictly applied. 8. Concluding Remarks There is no doubt that the multidisciplinary polymer reaction engineering horizons will be extended even further in the future as more value-added specialty polymers (especially those arising out of and applied to biological systems) are developed, and the material science field develops in this area. Certainly more effort is required in relating the fundamental properties of

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 1007

polymers, i.e., average molar mass and molar mass distribution, monomer sequence distribution, branching, etc., to particular requirements and attributes of specific applications. Developing a polymerization model is not an isolated effort; rather, it requires a commitment to the long-term development in parallel of a comprehensive polymerization database for characteristics of monomers, initiators, solvents, emulsifiers, chain-transfer agents, inhibitors, etc. Property databases, with the potential to provide an effective computer-based bridge between the simulation phase and detailed engineering, may have a lot to offer to these future directions. Acknowledgment M.A.D., A.P., and J.B.P.S. thank Professor Hamielec for being a great mentor. Financial support for this research supplied by the Natural Sciences and Engineering Research Council (NSERC) of Canada, the Ontario Centre for Materials Research (OCMR), ICI Worldwide, and Uniroyal Chemical Co. over the years is gratefully acknowledged. Nomenclature
a ) partition coefficient for monomer radicals between water phase and particles a ) root-mean-square end-to-end distance per square root of the number of monomer units A ) free-volume theory parameter a2 ) adjustable parameter Ad ) total surface area of monomer droplets (dm2) Aj ) jacket heat-transfer area (m2) Am ) total free micellar area (dm2) Ap ) total surface area of polymer particles (dm2) B ) free-volume theory parameter BN3 ) average number of trifunctional branch points (#) BN4 ) average number of tetrafunctional branch points (#) c ) mass concentration of accumulated polymer in reaction mixture (g L-1) C ) parameter modifying rate of change of initiator efficiency Cfcta ) transfer to chain-transfer agent constant Cfm ) transfer to monomer constant Cfmsi ) transfer to monomer-soluble impurities constant Cfp ) transfer to polymer constant Ck ) internal double-bond reaction constant Cpi ) heat capacity of species i (cal g-1 K-1) [CTA]p ) concentration of chain-transfer agent in the particles (mol L-1) D ) reaction diffusion coefficient dm ) density of monomer (kg L-1) dp ) particle diameter (dm) Dp ) diffusivity of monomer radicals in the particles (cm2 s-1) [drops] ) concentration of monomer droplets (mol L-1) Dw ) diffusivity of monomer radicals in the water phase (cm2 s-1) f, fo ) overall and initial initiator efficiency Fi,in ) flow of species i into the reactor (mol min-1) Fj ) instantaneous polymer composition (mole fraction monomer j bound in polymer) Fj ) cumulative polymer composition (mole fraction monoh mer j bound in copolymer) fj ) mole fraction of monomer j in the particles Fjw ) mole fraction of monomer j bound in the polymer formed in the water phase Hpi ) heat of polymerization of monomer i (cal g mol-1) [I] ) concentration of initiator (mol L-1) w jc ) entanglement spacing of pure polymer (# monomer units)

jcr ) overall critical chain length of polymer radical formed in the water phase (#) jcrj ) critical chain length of polymer radical formed in the water phase ending in species j (#) K ) terminal double-bond reaction constant k1 ) iron oxidation rate constant (L mol-1 min-1) k2 ) iron reduction rate constant (L mol-1 min-1) K3 ) free-volume theory parameter ka ) radical absorption by particles rate coefficient (dm min-1) Kc ) controller gain kcm ) capture by micelles rate constant (dm min-1) kcmd ) capture by monomer droplets rate constant (dm min-1) kcp ) capture by particles rate constant (dm min-1) Kcta ) partition coefficient for chain-transfer agent between the droplet and particle phases kd ) initiator decomposition rate constant (min-1) kdes ) radical desorption rate constant (min-1) kdes ) particle-size-independent radical desorption rate constant (min-1) kF ) particle coagulation rate constant (Lw min-1) kfcta ) overall rate constant for transfer to chain-transfer agent (L mol-1 min-1) kfctai ) rate constant for transfer to chain-transfer agent from radical ending in monomer i (L mol-1 min-1) kfidb, k** ) rate constant for reaction with internal double p bonds (L mol-1 min-1) kfm ) overall rate constant for transfer to monomer (L mol-1 min-1) kfmij ) rate constant for transfer to monomer j from radical ending in monomer i (L mol-1 min-1) kfmsi ) rate constant for transfer to monomer-soluble impurity (L mol-1 min-1) kfp ) overall rate constant for transfer to polymer (L mol-1 min-1) kfpij ) rate constant for transfer to monomer j on a dead polymer from monomer i (L mol-1 min-1) kftdb, k* ) rate constant for reaction with terminal double p bonds (L mol-1 min-1) kh, kho ) homogeneous nucleation rate constant (min-1) Kjmw ) partition coefficient for species j between the monomer droplet and water phases Kjwp ) partition coefficient for species j between the water and particle phases Kmsi ) partition coefficient for monomer-soluble impurity between the droplet and particle phases kp, kpo ) overall propagation rate constant (L mol-1 min-1) kp ) rate constant for reinitiation of oligomer radicals from monomer radicals (L mol-1 min-1) kpij ) propagation rate constant for radical ending in species i adding monomer j (L mol-1 min-1) kpIj ) propagation rate constant for addition of monomer j to an initiator radical (L mol-1 min-1) kt ) termination rate constant (L mol-1 min-1) kT ) translational diffusion-controlled termination rate constant (L mol-1 min-1) kto ) termination rate constant at zero polymer concentration (L mol-1 min-1) ktc ) termination by combination rate constant (L mol-1 min-1) ktd ) termination by disproportionation rate constant (L mol-1 min-1) h ktN, ktW, ktZ ) number-, weight-, and Z-average termination h h constants (L mol-1 min-1) ktr ) transfer rate constant (L mol-1 min-1) ktrd ) reaction diffusion-controlled termination rate constant (L mol-1 min-1) ktres ) residual termination rate constant (L mol-1 min-1) ktrm ) transfer to monomer rate constant (L mol-1 min-1) ktrt ) transfer to CTA rate constant (L mol-1 min-1)

1008

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 [R] ) concentration of initiator radicals (mol L-1) I w RI ) rate of initiation (mol L-1 min-1) w [Rj,i]w ) concentration of radicals of chain length j ending in monomer i in the water phase (mol L-1) w rmic ) radius of a micelle (dm) j rn, rn ) instantaneous and accumulated number-average chain lengths, respectively j rw, rw ) instantaneous and accumulated weight-average chain lengths, respectively Rn,i ) radical of chain length n ending in monomer i Rp ) overall rate of polymerization (mol L-1 min-1) p Rpj ) rate of polymerization of monomer j (mol L-1 min-1) p Rpjp ) rate of polymerization of monomer j in the particles (mol L-1 min-1) p Rpjw ) rate of polymerization of monomer j in the water phase (mol L-1 min-1) w Rtc, Rtd ) rate of termination by combination and disproportionation, respectively (mol L-1 min-1) [RTOT]des ) concentration of desorbed radicals in the water phase (mol L-1) w [RTOT]w, [R]w ) total concentration of radicals in the water phase (mol L-1) w [RTOT]wdrop, [R ]wmic, [R ]wpar ) concentration of radiTOT TOT cals in the water phase able to be captured by droplets, micelles, and particles, respectively (mol L-1) w [RTOT]wdrop ) concentration of radicals in the water phase that may undergo homogeneous nucleation (mol L-1) w Sa ) surface area covered by one molecule of emulsifier i 2 molecule-1) (dm [S]CMC ) critical micelle concentration (mol L-1) w [S]t ) total concentration of emulsifier (mol L-1) w t ) time (min) T, Tj, Tref, Tset ) reaction, jacket, reference, and setpoint temperatures, respectively (K) Tgi ) glass transition temperature of species i (K) Vaq ) volume of the aqueous phase (L) VF ) free volume (L) VFcrif, VFcrit, VFpcrit ) critical free volumes (L) Vi ) volume of species i (L) Vjmd, Vjmp, Vjmw ) volumes of monomer j in the droplets, aqueous phase, and particles, respectively (L) Vm ) volume of the monomer droplet phase (L) Vm ) molar volume (L mol-1) Vmj ) total volume of monomer j (L) Vo ) total volume of the organic (monomer and polymer) phase (L) vout ) total volumetric flow out of the reactor (L min-1) vp ) average volume of a polymer particle (Lp) Vp, V, Vo ) total volume of polymer particles (Lp) VT ) total volume of reaction mixture (L) Vw ) volume of water (L) W(r,y) ) weight fraction of polymer of chain length r and composition deviation y [WSI]i ) concentration of water-soluble impurity i (mol L-1) w x ) monomer conversion on a mass basis xc ) critical monomer conversion where droplets disappear Yo ) total concentration of radicals in the particles (mol L-1) p Greek Letters R ) grouping of terms representing radical entry into the polymer particles ) molecular weight contribution due to termination by combination ) reaction radius ) fraction of termination by disproportionation ) measurement error i ) chemical potential of species i

ktsb ) termination rate constant of polymer radicals in the sol with those on the gel (L mol-1 min-1) ktseg ) segmental diffusion-controlled termination rate constant (L mol-1 min-1) ktw ) water phase termination rate constant (L mol-1 min-1) kzj ) rate constant for reaction with water-soluble impurity j (L mol-1 min-1) L ) critical chain length for water phase polymerization lo ) length of a monomer unit m ) grouping of terms representing radical exit from the polymer particles m ) adjustable parameter for free-volume theory m, mi ) coefficient relating to volume characteristics in partition coefficient equations md ) partition coefficient for monomer radicals between the water phase and the particles Mj ) monomer of type j [Mj]m, [Mj]p, [Mj]w ) concentration of monomer j in the droplets, particles, and water phase, respectively (mol L-1) Mn ) accumulated number-average molecular weight h [M]p ) total concentration of monomer in the particles (mol L-1) p [M]psat ) total (saturation) concentration of monomer in the particles (mol L-1) p [MSI]p ) concentration of monomer-soluble impurities in the particles (mol L-1) p [M]w ) concentration of monomer in the water phase (mol L-1) w Mw ) accumulated weight-average molecular weight h Mwc ) saturation concentration of monomer in the water phase (mol L-1) w Mwcrit ) critical accumulated weight-average molecular h weight Mweff ) effective molecular weight of copolymer (g mol-1) MWi ) molecular weight of species i (g mol-1) Mwsat ) saturation concentration of monomer in the water phase (mol L-1) w Mwt ) molecular weight of monomer (g mol-1) Mz ) accumulated z-average molecular weight h n ) adjustable parameter for free-volume theory n ) average number of radicals per particle j NA ) Avogadros number (molecule mol-1) Nhom ) number of particles formed by homogeneous nucleation (# L-1) w Nj ) moles of species j (mol) Nmj ) moles of monomer j (mol) Nmic ) number of particles formed by micellar nucleation (# L-1) w Np, NT ) total number of particles in the reactor (# L-1) w Npolj ) moles of monomer j bound in the copolymer (mol) ns ) number of monomer units in one polymer chain segment Pm, Pn ) dead polymer molecule of chain length m or n Q0, Q1, Q2 ) zeroeth, first, and second moments of the molecular weight distribution (mol) Qj, QR ) heat losses in jacket and reactor, respectively (cal min-1) r ) particle radius (dm) r ) radial position in monomer-swollen particle (cm) R ) universal gas constant (cal mol-1 K-1) R ) radius of monomer-swollen particle (cm) rij ) reactivity ratio Rcm ) rate of capture of radicals by micelles (mol L-1 w min-1) Rcmd ) rate of capture of radicals by monomer droplets (mol L-1 min-1) w Rcp ) rate of capture of radicals by particles (mol L-1 w min-1)

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 1009 Fdes ) rate of radical desorption from the particles (mol L-1 min-1) w Fm ) density of monomer (kg L-1) Fp ) density of polymer (kg L-1) ) Lennard-Jones diameter ) molecular weight contribution due to termination by disproportionation and transfer to small molecules ) time (min) i ) mole fraction of radicals ending in monomer i in the particles p ) volume fraction of polymer in the particles i, ij ) Flory-Huggins interaction parameter
Bassett, D. R., Hamielec, A. E., Eds. Emulsion Polymers and Emulsion Polymerization. ACS Symposium Series; American Chemical Society: Washington, DC, 1981; Vol. 165. Behnken, D. W.; Horowitz, J.; Katz, S. Particle Growth Processes. Ind. Eng. Chem. Fundam. 1963, 2, 212. Benson, S. W.; North, A. M. The Kinetics of Free Radical Polymerization under Conditions of Diffusion-Controlled Termination. J. Am. Chem. Soc. 1962, 84, 935. Bhattacharya, D.; Hamielec, A. E. Bulk Thermal Copolymerization of Styrene/p-Methylstyrene: Modelling Diffusion-Controlled Termination and Propagation using Free-Volume Theory. Polymer 1986, 27, 611. Bohm, L. L. Reaction Model for Ziegler-Natta Polymerization Processes. Polymer 1978, 19, 545. Bonini, F.; Fraaije, V.; Fink, G. Propylene Polymerization Through Supported Metallocene/MAO Catalyst: Kinetic Analysis and Modelling. J. Polym. Sci., Part A: Polym. Chem. 1995, 33, 2393. Boor, J., Jr. Ziegler-Natta Catalysts and Polymerization; Academic Press: New York, 1979. Boots, H. J. M. Polymerization and Reptation, an Analytical Treatment. J. Polym. Sci., Polym. Phys. Ed. 1982, 20, 1695. Box, M. J. Improved Parameter Estimation. Technometrics 1970, 12, 219. Braun, D.; Bendlein, W.; Mott, G. A Simple Method of Determining Copolymerization Reactivity Ratios by Means of a Computer. Eur. Polym. J. 1973, 9, 1007. Britt, H. I.; Luecke, R. H. The Estimation of Parameters in Nonlinear, Implicit Models. Technometrics 1973, 15, 233. Broadhead, T. O.; Hamielec, A. E.; MacGregor, J. F. Dynamic Modelling of the Batch Semi-Batch and Continuous Production of Styrene/Butadiene Copolymers by Emulsion Polymerization. Makromol. Chem., Suppl. 1985, 10/11, 105. Brockmeier, N. F.; Rogan, J. B. Simulation of Continuous Propylene Polymerization in a Backmix Reactor Using Semibatch Kinetic Data. AIChE Symp. Ser. 1976, 72, 29. Brown, P. G.; Fujimori, K.; Brown, A. S.; Tucker, D. J. The Applicability of Copolymer Composition and Sequence Distribution Data to the Mechanistic Study of the Alternating Copolymerization of Maleic Anhydride with Substituted Styrene Derivatives. Makromol. Chem. 1993, 194, 1357. Buback, M.; Degener, B. Rate Coefficients for Free-Radical Polymerization of Butyl Acrylate to High Conversion. Makromol. Chem. 1993, 194, 2875. Buback, M.; Degener, B.; Huckestein, B. Conversion Dependence of Free-Radical Polymerization Rate Coefficients from LaserInduced Experiments, 1. Butyl Acrylate. Makromol. Chem., Rapid Commun. 1989, 10, 311. Buback, M.; Huckestein, B.; Russell, G. T. Modeling of Termination in Intermediate and High Conversion Free Radical Polymerizations. J. Macromol. Chem. Phys. 1994, 195, 539. Bukatov, G. D.; Zaikovskii, V. I.; Zakharov, V. A.; Kryukova, G. N.; Fenelonov, V. B.; Zagrafskaya, R. V. The Morphology of Polypropylene Granules and its Link with the Titanium Trichloride Texture. Polym. Sci. USSR 1982, 24, 599. Buls, V. W.; Higgins, T. L. A Particle Growth Theory for Heterogeneous Ziegler Polymerization. J. Polym. Sci., Polym. Chem. Ed. 1970, 8, 1037. Burke, A. L.; Duever, T. A.; Penlidis, A. Revisiting the Design of Experiments for Copolymer Reactivity Ratio Estimation. J. Polym. Sci., Part A: Polym. Chem. 1993, 31, 3065. Burke, A. L.; Duever, T. A.; Penlidis, A. Model Discrimination via Designed Experiments: Discriminating between the Terminal and Penultimate Models on the Basis of Composition Data. Macromolecules 1994a, 27, 386. Burke, A. L.; Duever, T. A.; Penlidis, A. Model Discrimination via Designed Experiments: Discriminating between the Terminal and Penultimate Models Based on Triad Fraction Data. Macromol. Theory Simul. 1994b, 3, 1005. Burke, A. L.; Duever, T. A.; Penlidis, A. Model Discrimination via Designed Experiments: Discrimination between the Terminal and Penultimate Models Based on Rate Data. Chem. Eng. Sci. 1995, 50, 1619. Burke, A. L.; Duever, T. A.; Penlidis, A. An Experimental Verification of Statistical Discrimination between the Terminal and Penultimate Copolymerization Models. J. Polym. Sci., Part A: Polym. Chem. 1996, 34, 2665. Canegallo, S.; Storti, G.; Morbidelli, M.; Carra, S. Densimetry for On-Line Conversion Monitoring in Emulsion Homo- and Copolymerization. J. Appl. Polym. Sci. 1993, 47, 961.

Literature Cited
Achilias, D. S.; Kiparissides, C. Development of a General Mathematical Framework for Modeling Diffusion-Controlled FreeRadical Polymerization Reactions. Macromolecules 1992, 25, 3739. Achilias, D. S.; Kiparissides, C. On the Validity of the SteadyState Approximations in High Conversion Diffusion-Controlled Free-Radical Copolymerization Reactions. Polymer 1994, 35, 1714. Adams, M. E.; Casey, B. S.; Mills, M. F.; Russell, G. T.; Napper, D. H.; Gilbert, R. G. High-Conversion Emulsion, Dispersion and Suspension Polymerization. Makromol. Chem., Macromol. Symp. 1990, 35/36, 1. Alfrey, T.; Goldfinger, G. Copolymerization of Systems of Three and More Components. J. Chem. Phys. 1944, 12, 322. Ali, S. I.; Zollars, R. L. Effects of Ionizable Groups on the Adsorption of Surfactants onto Latex Particle Surfaces. AIChE Annual Meeting, Chicago, Nov 1985; Paper 108g. Andersen, H. M.; Proctor, S. I. Redox Kinetics of the Peroxydisulfate-Iron-Sulfoxylate System. J. Polym. Sci., Part A 1965, 3, 2343. Anseth, K. S.; Bowman, C. N.; Peppas, N. A. Polymerization Kinetics and Volume Relaxation Behavior of Photopolymerized Multifunctional Monomers Producing Highly Cross-Linked Networks. J. Polym. Sci., Polym. Chem. Ed. 1994a, 32, 139. Anseth, K. S.; Wang, C. M.; Bowman, C. N. Kinetic Evidence of Reaction Diffusion during the Polymerization of Multi(meth)acrylate Monomers. Macromolecules 1994b, 27, 650. Armitage, P. D.; de la Cal, J. C.; Asua, J. M. Improved Methods for Solving Monomer Partitioning in Emulsion Copolymer Systems. J. Appl. Polym. Sci. 1994, 51, 1985. Arzamendi, G.; Asua, J. M. Monomer Addition Policies for Copolymer Composition Control in Semicontinuous Emulsion Copolymerization. J. Appl. Polym. Sci. 1989, 38, 2019. Arzamendi, G.; Asua, J. M. Copolymer Composition Control during the Seeded Emulsion Copolymerization of Vinyl Acetate and Methyl Acrylate. Makromol. Chem., Macromol. Symp. 1990, 35/ 36, 249. Arzamendi, G.; Asua, J. M. Copolymer Composition Control of Emulsion Copolymers in Reactors with Limited Capacity for Heat Removal. Ind. Eng. Chem. Res. 1991, 30, 1342. Arzamendi, G.; Leiza, J. R.; Asua, J. M. Semicontinuous Emulsion Copolymerization of Methyl Methacrylate and Ethyl Acrylate. J. Polym. Sci., Part A: Polym. Chem. 1991, 29, 1549. Arzamendi, G.; de la Cal, J. C.; Asua, J. M. Optimal Monomer Addition Policies for Composition Control of Emulsion Terpolymers. Angew. Makromol. Chem. 1992, 194, 47. Askill, I. N.; Gilding, D. K. A Monomer Feed System for Producing Comonomers of Widely Differing Reactivities. Polymer 1981, 22, 342. Asua, J. M.; Sudol, E. D.; El-Aasser, M. S. Radical Desorption in Emulsion Polymerization. J. Polym. Sci., Polym. Chem. 1989, 27, 3903. Ballard, M. J.; Gilbert, R. G.; Napper, D. H. Improved Methods for Solving the Smith-Ewart Equations in the Steady State. J. Polym. Sci., Polym. Lett. Ed. 1981a, 19, 533. Ballard, M. J.; Napper, D. H.; Gilbert, R. G. Theory of Emulsion Copolymerization Kinetics. J. Polym. Sci., Polym. Chem. Ed. 1981b, 19, 939. Ballard, M. J.; Gilbert, R. G.; Napper, D. H.; Pomery, P. J.; OSullivan, P. W.; ODonnell, J. H. Propagation Rate Coefficients from Electron Spin Resonance Studies of the Emulsion Polymerization of Methyl Methacrylate. Macromolecules 1986, 19, 1303.

1010 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997


Canegallo, S.; Canu, P.; Morbidelli, M.; Storti, G. Composition Control in Emulsion Copolymerization. II. Application to Binary and Ternary Systems. J. Appl. Polym. Sci. 1994, 54, 1919. Canu, P.; Canegallo, S.; Morbidelli, M.; Storti, G. Composition Control in Emulsion Copolymerization. I. Optimal Monomer Feed Policies. J. Appl. Polym. Sci. 1994, 54, 1899. Cardenas, J.; ODriscoll, K. F. High-Conversion Polymerization. I. Theory and Application to Methyl Methacrylate. J. Polym. Sci., Polym. Chem. Ed. 1976, 14, 883. Casey, B. S.; Morrison, B. R.; Maxwell, I. A.; Gilbert, R. G.; Napper, D. H. Free Radical Exit in Emulsion Polymerization. I. Theoretical Model. J. Polym. Sci., Part A: Polym. Chem. 1994, 32, 605. Charmot, D.; Guillot, J. Kinetic Modelling of Network Formation in Styrene-Butadiene Emulsion Copolymers: a Comparative Study with the Generalized Form of Florys Theory of Gelation. Polymer 1992, 33, 352. Cheng, H. N.; Kakugo, M. 13C NMR Analysis of Compositional Heterogeneity in Ethylene-Propylene Copolymers. Macromolecules 1991, 24, 1724. Chern, C. S.; Poehlein, G. W. Polymerization in Nonuniform Latex Particles. II. Kinetics of Two-Phase Emulsion Polymerization. J. Polym. Sci., Part A: Polym. Chem. 1990, 28, 3055. Chern, C. S.; Kuo, Y. N. Shear-Induced Coagulation Kinetics of Semibatch Seeded Emulsion Polymerization. Chem. Eng. Sci. 1996, 51, 1079. Chien, D. C. H.; Penlidis, A. On-Line Sensors for Polymerization Reactors. J. Macromol. Sci., Rev. Macromol. Chem. Phys. 1990, C30, 1. Chien, D. C. H.; Penlidis, A. Effect of Impurities on Continuous Solution Methyl Methacrylate Polymerization Reactors. I. OpenLoop Process Identification Results. Polym. React. Eng. 1994a, 2, 163. Chien, D. C. H.; Penlidis, A. Effect of Impurities on Continuous Solution Methyl Methacrylate Polymerization Reactors. II. Closed-Loop Process Real-Time Control. Chem. Eng. Sci. 1994b, 49, 1855. Chien, J. C.; Kuo, C.; Ang, T. J. Magnesium Chloride Supported High Mileage Catalyst for Olefin Polymerization. VI. Definitive Evidence Against Diffusion Limitation. J. Polym. Sci., Polym. Chem. Ed. 1985, 23, 723. Chiu, W. Y.; Carratt, G. M.; Soong, D. S. A Computer Model for the Gel Effect in Free-Radical Polymerization. Macromolecules 1983, 16, 348. Choi, K. Y.; Zhao, X.; Tang, S. Population Balance Modeling for a Continuous Gas Phase Olefin Polymerization Reactor. J. Appl. Polym. Sci. 1994, 53, 1589. Cooper, W. Kinetics of Polymerization Initiated by Ziegler-Natta and Related Catalyst. In Comprehensive Chemical Kinetics; Bamford, C. H.; Tipper, C. F. H., Eds.; Elsevier: Amsterdam, The Netherlands, 1976. Coyle, D. J.; Tulig, T. J.; Tirrell, M. Finite Element Analysis of High Conversion Free-Radical Polymerization. Ind. Eng. Chem. Fundam. 1985, 24, 343. Cozewith, C.; VerStrate, G. Ethylene-Propylene Copolymers. Reactivity Ratios, Evaluation and Significance. Macromolecules 1971, 4, 482. Crabtree, J. R.; Grimsby, F. N.; Nummelin, A. J.; Sketchley, J. M. The Role of Diffusion in the Ziegler Polymerization of Ethylene. J. Appl. Polym. Sci. 1973, 17, 959. Davis, T. P.; ODriscoll, K. F.; Piton, M. C.; Winnik, M. A. Determination of Propagation Rate Constants for the Copolymerization of Methyl Methacrylate and Styrene Using a Pulsed Laser Technique. J. Polym. Sci., Polym. Lett. 1989, C27, 181. de Carvalho, A. B.; Gloor, P. E.; Hamielec, A. E. A Kinetic Mathematical Model for Heterogeneous Ziegler-Natta Copolymerization. Polymer 1989, 30, 280. de Carvalho, A. B.; Gloor, P. E.; Hamielec, A. E. A Kinetic Mathematical Model for Heterogeneous Ziegler-Natta Copolymerization. Part 2. Stereochemical Sequence Length Distribution. Polymer 1990, 31, 1294. de la Cal, J. C.; Urzay, R.; Zamora, A.; Forcada, J.; Asua, J. M. Simulation of the Latex Particle Morphology. J. Polym. Sci., Part A: Polym. Chem. 1990, 28, 1011. de la Cal, J. C.; Echevarria, A.; Meira, G. R.; Asua, J. M. MinimumTime Strategy to Produce Nonuniform Emulsion Copolymers. I. Theory. J. Appl. Polym. Sci. 1995, 57, 1063. de la Rosa, L. V.; Sudol, E. D.; El-Aasser, M. S.; Klein, A. Details of the Emulsion Polymerization of Styrene Using a Reaction Calorimeter. J. Polym. Sci., Part A: Polym. Chem. 1996, 34, 461. Dimitratos, J.; Georgakis, C.; El-Aasser, M. S.; Klein, A. Dynamic Modeling and State Estimation for an Emulsion Copolymerization Reactor. Comput. Chem. Eng. 1989, 13, 21. Dougherty, E. P. The SCOPE Dynamic Model for Emulsion Polymerization I. Theory. J. Appl. Polym. Sci. 1986, 32, 3051. Dube, M. A. A Systematic Approach to the Study of Multicompo nent Polymerization Kinetics. Ph.D. Thesis, University of Waterloo, Waterloo, Ontario, Canada, 1994. Dube, M. A.; Penlidis, A. A Systematic Approach to the Study of Multicomponent Polymerization Kinetics: The Butyl Acrylate/ Methyl Methacrylate/Vinyl Acetate Example I. Bulk Copolymerization. Polymer 1995a, 36, 587. Dube, M. A.; Penlidis, A. A Systematic Approach to the Study of Multicomponent Polymerization Kinetics: The Butyl Acrylate/ Methyl Methacrylate/Vinyl Acetate Example 2. Bulk (and Solution) Terpolymerization. Macromol. Chem. Phys. 1995b, 196, 1101. Dube, M. A.; Penlidis, A. A Systematic Approach to the Study of Multicomponent Polymerization Kinetics: The Butyl Acrylate/ Methyl Methacrylate/Vinyl Acetate Example 3. Emulsion Homoand Copolymerization in a Pilot Plant Reactor. Polym. Int. 1995c, 37, 235. Dube, M. A.; Penlidis, A. Emulsion Terpolymerization of Butyl Acrylate/Methyl Methacrylate/Vinyl Acetate: Experimental Results. J. Polym. Sci., Polym. Chem. 1997, in press. Dube, M. A.; Amin Sanayei, R.; Penlidis, A.; ODriscoll, K. F.; Reilly, P. M. A Microcomputer Program for Estimation of Copolymerization Reactivity Ratios. J. Polym. Sci., Part A: Polym. Chem. 1991a, 29, 703. Dube, M. A.; Rilling, K.; Penlidis, A. A Kinetic Investigation of Butyl Acrylate Polymerization. J. Appl. Polym. Sci. 1991b, 43, 2137. Duerksen, J. H. Polymerization of Styrene in Continuous Stirred Tank Reactors. Ph.D. Thesis, McMaster University, Hamilton, Ontario, Canada, 1968. Duerksen, J. H.; Hamielec, A. E. Polymer Reactors and Molecular Weight Distributions. Part I: Free-Radical Polymerization in a Continuous Stirred Tank Reactor. AIChE J. 1967, 13, 1081. Duerksen, J. H.; Hamielec, A. E. Polymer Reactors and Molecular Weight Distributions. IV. Free-Radical Polymerization in a Steady-State Stirred-Tank Reactor Train. J. Polym. Sci., Part C 1968, 25, 155. Duever, T. A.; ODriscoll, K. F.; Reilly, P. M. The Use of the Errorin-Variables Model in Terpolymerization. J. Polym. Sci., Polym. Chem. 1983, 21, 2003. Dunn, A. S. Rate of emulsifier adsorption as a factor in the nucleation of polymer latex particles. In ACS Symposium Series; Daniels, E. S., Sudol, E. D., El-Aasser, M. S., Eds.; American Chemical Society: Washington, DC, 1992; Vol. 492, p 45. Echevarria, A.; de la Cal, J. C.; Asua, J. M. Minimum-Time Strategy to Produce Nonuniform Emulsion Copolymers. II. Open-Loop Control. J. Appl. Polym. Sci. 1995, 57, 1217. El-Aasser, M. S.; Makgawinata, T.; Misra, S.; Vanderhoff, J. W.; Pichot, C.; Llauro, M. F. Preparation, Characterization and Properties of Vinyl Acetate-Butyl Acrylate Copolymer Latexes. In Emulsion Polymerization of Vinyl Acetate; El-Aasser, M. S., Vanderhoff, J. W., Eds.; Applied Science Publishers: Oxford, U.K., 1981. El-Aasser, M. S.; Makgawinata, T.; Vanderhoff, J. W.; Pichot, C. Batch and Semicontinuous Emulsion Copolymerization of Vinyl Acetate-Butyl Acrylate. I. Bulk, Surface, and Colloidal Properties of Copolymer Latexes. J. Polym. Sci., Polym. Chem. Ed. 1983, 21, 2363. Elias, H. G. Plastics, General Survey. In Ullmanns Encyclopedia of Industrial Chemistry; Elvers, B., Hawkins, S., Schultz, G., Eds.; VCH Publishers: New York, 1992; Vol. A20, p 543. Emelie, B.; Pichot, C.; Guillot, J. Batch Emulsion Copolymerization of Butyl Acrylate and Methyl Methacrylate in the Presence of Sodium Dodecyl Sulfate. Makromol. Chem. 1991, 192, 1629. Ferrero, M. A.; Chiovetta, M. G. Catalyst Fragmentation during Polymerization: Part I. The Effects of Grain Size and Structure. Polym. Eng. Sci. 1987a, 27, 1436. Ferrero, M. A.; Chiovetta, M. G. Catalyst Fragmentation during Polymerization: Part II. Microparticle Diffusion and Reaction Effects. Polym. Eng. Sci. 1987b, 27, 1448.

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 1011


Ferrero, M. A.; Chiovetta, M. G. Catalyst Fragmentation during Polymerization. III: Bulk Polymerization Process Simulation. Polym. Eng. Sci. 1991, 31, 886. Fineman, W.; Ross, S. D. Linear Method for Determining Monomer Reactivity Ratios in Copolymerization. J. Polym. Sci. 1950, 5, 259. Fitch, R. M. Latex particle nucleation and growth. In ACS Symposium Series; Bassett, D. R., Hamielec, A. E., Eds.; American Chemical Society: Washington, DC, 1981; Vol. 165, p 1. Fitch, R. M.; Tsai, C. H. Polymer Colloids: Particle Formation in Nonmicellar Systems. Polym. Lett. 1970, 8, 703. Fitch, R. M.; Tsai, C. H. Particle Formation in Polymer Colloids. III: Prediction of the Number of Particles by a Homogeneous Nucleation Theory. In Polymer Colloids; Fitch, R. M., Ed.; Plenum Press: New York, 1971. Floyd, S.; Choi, K. Y.; Taylor, T. W.; Ray, W. H. Polymerization of Olefins through Heterogeneous Catalysis. IV. Modeling of Heat and Mass Transfer Resistance in the Polymer Particle Boundary. J. Appl. Polym. Sci. 1986a, 31, 2231. Floyd, S.; Choi, K. Y.; Taylor, T. W.; Ray, W. H. Polymerization of Olefins through Heterogeneous Catalysis. III. Polymer Particle Modelling with an Analysis of Intraparticle Heat and Mass Transport. J. Appl. Polym. Sci. 1986b, 31, 2935. Floyd, S.; Hutchinson, R. A.; Ray, W. H. Polymerization of Olefins through Heterogeneous Catalysis. V. Gas-Liquid Mass Transfer Limitations in Liquid Slurry Reactors. J. Appl. Polym. Sci. 1986c, 32, 5451. Floyd, S.; Heiskanen, T.; Ray, W. H. Solid Catalyzed Olefin Polymerization. Chem. Eng. Prog. 1988, 84, 56. Fontenot, K.; Schork, F. J. Simulation of Mini/Macro Emulsion Polymerizations. I. Development of the Model. Polym. React. Eng. J. 1992-93a, 1, 75. Fontenot, K.; Schork, F. J. Simulation of Mini/Macro Emulsion Polymerizations. II. Sensitivities and Experimental Comparison. Polym. React. Eng. J. 1992-93b, 1, 289. Friis, N.; Nyhagen, L. A Kinetic Study of the Emulsion Polymerization of Vinyl Acetate. J. Appl. Polym. Sci. 1973, 17, 2311. Friis, N.; Hamielec, A. E. Chapter in Intensive Short Course on Polymer Production Technology, 1977; available from A. E. Hamielec or A. Penlidis. Friis, N.; Goosney, D.; Wright, J. D.; Hamielec, A. E. Molecular Weight and Branching Development in Vinyl Acetate Emulsion Polymerization. J. Appl. Polym. Sci. 1974, 18, 1247. Froment, G. F.; Bischoff, K. B. Chemical Reactor Analysis and Design; John Wiley & Sons: New York, 1990. Fukuda, T.; Ma, Y. D.; Inagaki, H. Free Radical Copolymerization. 3. Determination of Rate Constants of Propagation and Termination for the Styrene/Methyl Methacrylate System. A Critical Test of Terminal Model Kinetics. Macromolecules 1985, 18, 17. Fukuda, T.; Ma, Y. D.; Inagaki, H. Free Radical Copolymerization. 6. New Interpretation for the Propagation Rate Versus Composition Curve. Makromol. Chem., Rapid Commun. 1987, 8, 495. Fukuda, T.; Ma, Y. D.; Kubo, K.; Takada, A. Free Radical Copolymerization. 7. Reinterpretation of Velocity of Copolymerization Data. Polym. J. 1989, 21, 1003. Fukuda, T.; Ma, Y. D.; Kubo, K.; Inagaki, H. Penultimate-Unit Effects in Free-Radical Copolymerization. Macromolecules 1991, 24, 370. Galli, P.; Haylock, J. C. Advances in Ziegler-Natta PolymerizationsUnique Polyolefin Copolymers, Alloys and Blends Made Directly in the Reactor. Macromol. Chem., Macromol. Symp. 1992, 63, 19. Galvan, R.; Tirrell, M. Orthogonal Collocation Applied to Analysis of Heterogeneous Ziegler-Natta Polymerization. Comput. Chem. Eng. 1986a, 10, 77. Galvan, R.; Tirrell, M. Molecular Weight Distribution Predictions for Heterogeneous Ziegler-Natta Polymerization Using a TwoSite Model. Chem. Eng. Sci. 1986b, 41, 2385. Gao, J.; Penlidis, A. A Comprehensive Simulator/Database Package for Reviewing Free-Radical Homopolymerizations. J. Macromol. Sci., Rev. Macromol. Chem. 1996, C36, 199. Garcia-Rubio, L. H.; Mehta, J. Initiation reactions and the modeling of polymerization kinetics. In ACS Symposium Series; Provder, T., Ed.; American Chemical Society: Washington, DC, 1986; Vol. 313, p 202. Garcia-Rubio, L. H.; Lord, M. G.; MacGregor, J. F.; Hamielec, A. E. Bulk Copolymerization of Styrene and Acrylonitrile: Experimental Kinetics and Mathematical Modelling. Polymer 1985, 26, 2001. Gugliotta, L. M.; Arzamendi, G.; Asua, J. M. Choice of Monomer Partition Model in Mathematical Modeling of Emulsion Copolymerization Systems. J. Appl. Polym. Sci. 1995a, 55, 1017. Gugliotta, L. M.; Leiza, J. R.; Arotena, M.; Armitage, P. D.; Asua, J. M. Copolymer Composition Control in Unseeded Emulsion Polymerization Using Calorimetric Data. Ind. Eng. Chem. Res. 1995b, 34, 3899. Guillot, J. Some Thermodynamic Aspects in Emulsion Copolymerization. Makromol. Chem. Suppl. 1985, 10/11, 235. Guyot, A.; Guillot, J.; Pichot, C.; Guerrero, L. R. New design for producing constant composition copolymers in emulsion polymerizationsComparison with other processes. In ACS Symposium Series; Bassett, D. R., Hamielec, A. E., Eds.; American Chemical Society: Washington, DC, 1981; Vol. 165, p 415. Guyot, A.; Guillot, J.; Graillat, C.; Llauro, M. F. Controlled Composition in Emulsion Copolymerization Application to Butadiene-Acrylonitrile Copolymers. J. Macromol. Sci., Chem. 1984, A21, 683. Hamielec, A. E. Synthesis Kinetics and Characterization of Poly(Vinyl Acetate)sMolecular Weight and Long Chain Branching Development. In Emulsion Polymerization of Vinyl Acetate; ElAasser, M. S., Vanderhoff, J. W., Eds.; Applied Science Publishers: Oxford, U.K., 1981. Hamielec, A. E.; Soares, J. B. P. Polymer Reactor Engineerings Metallocene Catalysis. Prog. Polym. Sci. 1996, 21, 651. Hamielec, A. E.; MacGregor, J. F.; Penlidis, A. Multicomponent Free-Radical Polymerization in Batch, Semi-Batch and Continuous Reactors. Makromol. Chem., Macromol. Symp. 1987, 10/11, 521. Hamielec, A. E.; MacGregor, J. F.; Penlidis, A. Copolymerization. In Comprehensive Polymer Science; Allen, Sir G., Ed.; Pergamon Press: Oxford, U.K., 1989; Vol. 3, p 17. Hansen, F. K. Is there life beyond micelles? Mechanisms of latex particle nucleation. In ACS Symposium Series; Daniels, E. S., Sudol, E. D., El-Aasser, M. S., Eds.; American Chemical Society: Washington, DC, 1992; Vol. 492, p 12. Hansen, F. K. The Function of Surfactant Micelles in Latex Particle Nucleation. Chem. Eng. Sci. 1993, 48, 437. Hansen, F. K.; Ugelstad, J. Particle Nucleation in Emulsion Polymerization. I. A Theory for Homogeneous Nucleation. J. Polym. Sci., Polym. Chem. Ed. 1978, 16, 1953. Hansen, F. K.; Ugelstad, J. Particle Nucleation in Emulsion Polymerization. IV. Nucleation in Monomer Droplets. J. Polym. Sci., Polym. Chem. Ed. 1979, 17, 3069. Hansen, F. K.; Ugelstad, J. Particle Formation Mechanisms. In Emulsion Polymerization; Piirma, I., Ed.; Academic Press: New York, 1982. Harada, M.; Nomura, M.; Eguchi, W.; Nagata, S. Studies of the Effect of Polymer Particles on Emulsion Polymerization. J. Chem. Eng. Jpn. 1971, 4, 54. Harwood, H. J. Structures and Compositions of Copolymers. Makromol. Chem., Macromol. Symp. 1987, 10/11, 331. Hendy, B. N. Preparation of Acrylonitrile-Styrene Copolymers by Calorimetrically Controlled Monomer Feeding. Adv. Chem. Ser. 1975, 142, 115. Herrmann, H. F.; Bohm, L. L. Particle Forming Process in Slurry Polymerization of Ethylene with Homogeneous Catalysts. Polym. Commun. 1991, 32, 58. Hill, D. J. T.; ODonnell, J. J.; OSullivan, P. W. Analysis of the Mechanism of Copolymerization of Styrene and Acrylonitrile. Macromolecules 1982, 15, 960. Hill, D. J. T.; ODonnell, J. J.; OSullivan, P. W. Methyl Methacrylate/Chloroprene Copolymerization: An Evaluation of Copolymerization Models. Polymer 1984, 25, 569. Hill, D. J. T.; Lang, A. P.; ODonnell, J. J.; OSullivan, P. W. Determination of Reactivity Ratios from Analysis of Triad FractionssAnalysis of the Copolymerization of Styrene-Acrylonitrile as Evidence for the Penultimate Model. Eur. Polym. J. 1989, 25, 911. Hock, C. W. How TiCl3 Catalysts Control the Texture of AsPolymerized Polypropylene. J. Polym. Sci., Polym. Chem. Ed. 1966, 4, 3055. Hoel, E. L.; Cozewith, C.; Byrne, G. D. Effect of Diffusion on Heterogeneous Ethylene Propylene Copolymerization. AIChE J. 1994, 40, 1669.

1012 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997


Hoffman, T. W. Chapter in Intensive Short Course on Polymer Production Technology; 1981; available from A. E. Hamielec or A. Penlidis. Hoffman, E. J. Kinetic Modelling of Emulsion Polymerization of Styrene and Acrylonitrile. M.Eng. Thesis, McMaster University, Hamilton, Ontario, Canada, 1984. Huang, J.; Rempel, G. L. Ziegler-Natta Catalysts for Olefin Polymerization: Mechanistic Insights from Metallocene Systems. Prog. Polym. Sci. 1995, 20, 459. Hui, A.; Hamielec, A. E. Polymer Reactors and Molecular Weight Distribution. V. Free-Radical Polymerization in a Transient Stirred-Tank Reactor Train. J. Polym. Sci., Part C 1968, 25, 167. Huo, B. P.; Campbell, J. D.; Penlidis, A.; MacGregor, J. F.; Hamielec, A. E. Effect of Impurities on Emulsion Polymerization: Case II Kinetics. J. Appl. Polym. Sci. 1988, 35, 2009. Hutchinson, R. A. Modeling of Free-Radical Polymerization Kinetics with Crosslinking for Methyl Methacrylate/Ethylene Glycol Dimethacrylate. Polym. React. Eng. 1992-93, 4, 521. Hutchinson, R. A.; Ray, W. H. Sorption Effects in Heterogeneous Catalyzed Olefin Polymerization. AIChE Annual Meeting, Washington, DC, Nov 1988; Paper 109a. Hutchinson, R. A.; Chen, C. M.; Ray, W. H. Polymerization of Olefins through Heterogeneous Catalysis. X: Modeling Particle Growth and Morphology. J. Appl. Polym. Sci. 1992, 44, 1389. Ito, K. Estimation of Molecular Weight in Terms of the Gel Effect in Radical Polymerization. Polym. J. 1980, 12, 499. Ito, K. Evaluation of Termination Rate by the Free Volume Theory in Radical Polymerization. Polym. J. 1981, 13, 727. Jones, K. M.; Bhattacharya, D.; Brash, J. L.; Hamielec, A. E. An Investigation of the Kinetics of Copolymerization of Methyl Methacrylate/p-Methyl Styrene to High Conversion: Modelling Diffusion-Controlled Termination and Propagation by FreeVolume Theory. Polymer 1986, 27, 602. Kakugo, M.; Sedatoshi, H.; Sakai, J.; Yokoyama, M. Growth of Polypropylene Particles in Heterogeneous Ziegler-Natta Polymerization. Macromolecules 1989, 22, 3172. Kaminsky, W. Polymerization and Copolymerization with a Highly Active, Soluble Ziegler-Natta Catalyst. In Catalytic Polymerization of Olefins; Keii, T., Soga, K., Eds.; Kodansha-Elsevier: Tokyo, 1986a. Kaminsky, W. Preparation of Special Polyolefins from Soluble Zirconium Compounds with Aluminoxane as Cocatalyst. In Catalytic Polymerization of Olefins; Keii, T., Soga, K., Eds.; Kodansha-Elsevier: Tokyo, 1986b. Kaminsky, W. Polymerization and Copolymerization of Olefins with Metallocene/Aluminoxane Catalysts. Catal. Soc. Jpn. 1991, 33, 536. Keii, T. Kinetics of Ziegler-Natta Polymerization; Kodansha: Tokyo, 1972. Keii, T. Propene Polymerization with a Magnesium ChlorideSupported Ziegler Catalyst, 1sPrincipal Kinetics. Makromol. Chem. 1982, 183, 2285. Keii, T.; Doi, Y.; Suzuki, E.; Tamura, M.; Murata, M.; Soga, K. Propene Polymerization with a Magnesium Chloride Supported Ziegler Catalyst. Makromol. Chem. 1984, 185, 1537. Kelen, T.; Tudos, F. Analysis of the Linear Methods for Determin ing Copolymerization Reactivity Ratios. 1. A New Improved Linear Graphic Method. J. Macromol. Sci., Chem. 1975, A9, 1. Kiparissides, C.; Ponnuswamy, S. R. Application of Population Balance Equations to Latex Reactors. Chem. Eng. Commun. 1981, 10, 283. Kshirsagar, R. S.; Poehlein, G. W. Radical Entry into Particles during Emulsion Polymerization of Vinyl Acetate. J. Appl. Polym. Sci. 1994, 54, 909. Lai, S. Y.; Wilson, J. R.; Knight, G. W.; Stevens, J. C.; Chum, P. W. S. Elastic substantially linear olefin polymers. U. S. Patent 5,272,236, 1993a. Lai, S. Y.; Wilson, J. R.; Knight, G. W.; Stevens, J. C. Elastic substantially linear olefin polymers. U. S. Patent Application WO 93/08221, 1993b. Laurence, R. L.; Chiovetta, M. G. Heat and Mass Transfer during Olefin Polymerization from the Gas Phase. In Polymer Reaction Engineering, Reichert, K. H., Geiseler, W., Eds.; Carl Hanser Verlag: Munich, 1983. Laurier, G. C.; ODriscoll, K. F.; Reilly, P. M. Estimating Reactivity in Free Radical Copolymerizations. J. Polym. Sci., Polym. Symp. 1985, 72, 17. Lee, T. Y.; Nitirahardjo, S.; Lee, S. An Analytic Approach in Kinetic Modelling of Ziegler-Natta Polymerization of Butadiene. J. Appl. Polym. Sci. 1994, 53, 1605. Leiza, J. R.; Arzamendi, G.; Asua, J. M. Copolymer Composition Control in Emulsion Polymerization Using Technical Grade Monomers. Polym. Int. 1993, 30, 455. Lepizzera, S. M.; Hamielec, A. E. Nucleation of Particles in Seeded Emulsion Polymerization of Vinyl Acetate with Poly(Vinyl Alcohol) as Emulsifier. Macromol. Chem. Phys. 1994, 195, 3103. Li, B.; Brooks, B. W. Prediction of the Average Number of Radicals per Particle for Emulsion Polymerization. J. Polym. Sci., Part A: Polym. Chem. 1993, 31, 2397. Litt, M. Kinetics of Styrene Emulsion Polymerization with a Stable Monomer Radical. Polym. Int. 1993, 30, 213. Lorenzini, P.; Bertrand, P.; Villermaux, J. Modelisation de la Copolymerisation Ethylene R-Olefine par Catalyse Ziegler ` Natta. Can. J. Chem. Eng. 1991, 69, 682. Lovell, P. A.; Shah, T. H.; Heatley, F. Chain Transfer to Polymer in Emulsion Polymerization of n-Butyl Acrylate Studied by 13C N.M.R. Spectroscopy and G.P.C. Polym. Commun. 1991, 32, 98. Lovell, P. A.; Shah, T. H.; Heatley, F. Correlation of the extent of chain transfer to polymer with reaction conditions for emulsion polymerization of n-butyl acrylate. In ACS Symposium Series; Daniels, E. S., Sudol, E. D., El-Aasser, M. S., Eds.; American Chemical Society: Washington, DC, 1992; Vol. 492, p 188. Lyons, R. A.; Hutovic, J.; Piton, M. C.; Christie, D. I.; Clay, P. A.; Manders, B. G.; Kable, S. H.; Gilbert, R. G. Pulsed-Laser Polymerization Measurements of the Propagation Rate Coefficient for Butyl Acrylate. Macromolecules 1996, 29, 1918. Mahabadi, H. K.; ODriscoll, K. F. Concentration Dependence of the Termination Rate Constant during the Initial Stages of Free Radical Polymerization. Macromolecules 1977, 10, 55. Mallya, P.; Plamthottam, S. S. Termination Rate Constant in Butyl Acrylate Batch Emulsion Polymerization. Polym. Bull. 1989, 21, 497. Marten, F. L.; Hamielec, A. E. High-conversion diffusion-controlled polymerization. In ACS Symposium Series; Henderson, H. N., Bouton, T. C., Eds.; American Chemical Society: Washington, DC, 1979; Vol. 104, p 43. Marten, F. L.; Hamielec, A. E. High-Conversion DiffusionControlled Polymerization of Styrene I. J. Appl. Polym. Sci. 1982, 27, 489. Maxwell, I. A.; Russell, G. T. Diffusion Controlled Copolymerization Kinetics. Makromol. Chem., Theory Simul. 1993, 2, 95. Maxwell, I. A.; Morrisson, B. R.; Napper, D. H.; Gilbert, R. G. Entry of Free Radicals into Latex Particles in Emulsion Polymerization. Macromolecules 1991, 24, 1629. Maxwell, I. A.; Kurja, J.; van Doremaele, G. H. J.; German, A. L. Thermodynamics of Swelling of Latex Particles with Two Monomers. Makromol. Chem. 1992a, 193, 2065. Maxwell, I. A.; Morrisson, B. R.; Napper, D. H.; Gilbert, R. G. The Effect of Chain Transfer Agent on the Entry of Free Radicals in Emulsion Polymerization. Makromol. Chem. 1992b, 193, 303. Mayo, F. R.; Lewis, F. M. Copolymerization. I. A Basis for Comparing the Behavior of Monomers in Copolymerization; the Copolymerization of Styrene and Methyl Methacrylate. J. Am. Chem. Soc. 1944, 66, 1594. McAuley, K. B.; MacGregor, J. F.; Hamielec, A. E. A Kinetic Model for Industrial Gas-Phase Ethylene Copolymerization. AIChE J. 1990, 36, 850. McFarlane, R. C.; Reilly, P. M.; ODriscoll, K. F. Comparison of the Precision of Estimation of Copolymerization Reactivity Ratios by Current Methods. J. Polym. Sci., Polym. Chem. Ed. 1980, 18, 251. McManus, N. T.; Penlidis, A. A Kinetic Investigation of Styrene/ Ethyl Acrylate Copolymerization. J. Polym. Sci., Part A: Polym. Chem. 1996, 34, 237. Mead, R. N.; Poehlein, G. W. Emulsion Copolymerization of Styrene-Methyl Acrylate and Styrene-Acrylonitrile in Continuous Stirred Tank Reactors. 1. Ind. Eng. Chem. Res. 1988, 27, 2283. Mead, R. N.; Poehlein, G. W. Emulsion Copolymerization of Styrene-Methyl Acrylate and Styrene-Acrylonitrile in Continuous Stirred Tank Reactors. 2. Aqueous-Phase Polymerization and Radical Capture. Ind. Eng. Chem. Res. 1989a, 28, 51. Mead, R. N.; Poehlein, G. W. Free-Radical Transport from Latex Particles. J. Appl. Polym. Sci. 1989b, 38, 105. Merz, E.; Alfrey, T.; Goldfinger, G. Intramolecular Reactions in Vinyl Polymers as a Means of Investigation of the Propagation Step. J. Polym. Sci. 1946, 1, 75.

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 1013


Min, K. W.; Ray, W. H. On the Mathematical Modeling of Emulsion Polymerization Reactors. J. Macromol. Sci., Rev. Macromol. Chem. 1974, C11, 177. Morbidelli, M.; Storti, G.; Carra, S. Role of Micellar Equilibria on Modelling of Batch Emulsion Polymerization Reactors. J. Appl. Polym. Sci. 1983, 28, 901. Moritz, H. U. Polymerization CalorimetrysA Powerful Tool for Reactor Control. In Polymer Reaction Engineering; Reichert, K. H., Geiseler, W., Eds.; VCH Verlag: Weinheim, Germany, 1989. Nagel, E. J.; Kirilov, V. A.; Ray, W. H. Prediction of Molecular Weight Distributions for High-Density Polyolefins. Ind. Eng. Chem., Product Res. Dev. 1980, 19, 372. Noel, L. F. J.; Van Zon, J. M. A. M.; Maxwell, I. A.; German, A. L. Prediction of Polymer Composition in Batch Emulsion Copolymerization. J. Polym. Sci., Part A: Polym. Chem. 1994, 32, 1009. Nomura, M. Desorption and Reabsorption of Free Radicals in Emulsion Polymerization. In Emulsion Polymerization; Piirma, I., Ed.; Academic Press: New York, 1982. Nomura, M.; Harada, M. Rate Coefficient for Radical Desorption in Emulsion Polymerization. J. Appl. Polym. Sci. 1981, 26, 17. Nomura, M.; Harada, M.; Nakagawara, K.; Eguchi, W.; Nagata, S. The Role of Polymer Particles in the Emulsion Polymerization of Vinyl Acetate. J. Chem. Eng. Jpn. 1971a, 4, 160. Nomura, M.; Kojima, H.; Harada, M.; Eguchi, W.; Nagata, S. Continuous Flow Operation in Emulsion Polymerization of Styrene. J. Appl. Polym. Sci. 1971b, 15, 675. Nomura, M.; Harada, M.; Eguchi, W.; Nagata, S. Kinetics and mechanism of the emulsion polymerization of vinyl acetate. In ACS Symposium Series; Piirma, I., Gardon, J. L., Eds.; American Chemical Society: Washington, DC, 1976; Vol. 24, p 102. Nomura, M.; Minamino, Y.; Fujita, K.; Harada, M. The Role of Chain Transfer Agents in the Emulsion Polymerization of Styrene. J. Polym. Sci., Polym. Chem. Ed. 1982, 20, 1261. Noristi, L.; Marchetti, E.; Baruzzi, G.; Sgarzi, P. Investigation on the Particle Growth Mechanism in Propylene Polymerization with MgCl2-Supported Ziegler-Natta Catalysts. J. Polym. Sci., Part A: Polym. Chem. 1994, 32, 3047. ODriscoll, K. F.; Reilly, P. M. Determination of Reactivity Ratios in Copolymerization. Makromol. Chem., Makromol. Symp. 1987, 10/11, 355. Okubo, M.; Fujimura, M.; Mori, T. The Acceleration of Decomposition of Potassium Persulfate in the Presence of Sodium Dodecyl Sulfate and Polymer Particles as a Model of Emulsion Polymerization System. Colloid Polym. Sci. 1991, 269, 121. Olaj, O. F.; Zifferer, G.; Gleixner, G. Termination Processes in Free Radical Polymerization. 8. Complete Treatment of a Kinetic Scheme Comprising Chain Length Dependent Termination in Terms of Closed and Approximate Closed Expressions Based on the Geometric Mean Assumption. Macromolecules 1987, 20, 839. Omi, S.; Kushibiki, K.; Iso, M. The Computer Modeling of Multicomponent, Semibatch Emulsion Copolymerization. Polym. Eng. Sci. 1987, 27, 470. OShaughnessy, B.; Yu, J. Autoacceleration in Free Radical Polymerization. 1. Conversion. Macromolecules 1994a, 27, 5067. OShaughnessy, B.; Yu, J. Autoacceleration in Free Radical Polymerization. 2. Molecular Weight Distributions. Macromolecules 1994b, 27, 5079. OToole, J. T. Kinetics of Emulsion Polymerization. J. Appl. Polym. Sci. 1965, 9, 1291. Panke, D. Polymerization of Methyl Methacrylate up to High Degrees of Conversion: Model Calculations Considering the Presence of a Prepolymer. Makromol. Chem., Rapid Commun. 1986, 7, 171. Patino-Leal, H.; Reilly, P. M. A Bayesian Study of the Error-inVariables Model. Technometrics 1981, 23, 221. Patino-Leal, H.; Reilly, P. M.; ODriscoll, K. F. On the Estimation of Reactivity Ratios. J. Polym. Sci., Polym. Lett. 1980, 18, 219. Penlidis, A. Latex Production TechnologysReactor Design Considerations. Ph.D. Thesis, McMaster University, Hamilton, Ontario, Canada, 1986. Penlidis, A.; MacGregor, J. F.; Hamielec, A. E. A Theoretical and Experimental Investigation of the Batch Emulsion Polymerization of Vinyl Acetate. Polym. Proc. Eng. 1985a, 3, 185. Penlidis, A.; MacGregor, J. F.; Hamielec, A. E. Dynamic Modeling of Emulsion Polymerization Reactors. AIChE J. 1985b, 31, 881. Penlidis, A.; MacGregor, J. F.; Hamielec, A. E. Mathematical modelling of emulsion polymerization reactors: A population balance approach and its applications. In ACS Symposium Series; Provder, T., Ed.; American Chemical Society: Washington, DC, 1986; Vol. 313, p 219. Penlidis, A.; MacGregor, J. F.; Hamielec, A. E. Effect of Impurities on Emulsion Polymerization: Case I. Kinetics. J. Appl. Polym. Sci. 1988, 35, 2023. Penlidis, A.; Ponnuswamy, S. R.; Kiparissides, C.; ODriscoll, K. F. Polymer Reaction Engineering: Modelling Considerations for Control Studies. Chem. Eng. J. 1992, 50, 95. Poehlein, G. W. Personal communication, Georgia Institute of Technology, 1990. Poehlein, G. W.; Lee, H. C.; Chern, C. S. Free Radical Transport and Reactions in Emulsion Polymerization. In Polymer Reaction Engineering; Reichert, K. H., Geiseler, W., Eds.; Huthig and Wepf: Berlin, 1986. Polymer Handbook; Brandrup, J., Immergut, E. H., Eds.; Wiley: New York, 1989. Rawlings, J. B.; Ray, W. H. The Modeling of Batch and Continuous Emulsion Polymerization Reactors. Part 1: Model Formulation and Sensitivity to Parameters. Polym. Eng. Sci. 1988, 28, 237. Ray, W. H. On the Mathematical Modeling of Polymerization Reactors. J. Macromol. Sci., Rev. Macromol. Chem. 1972, C8, 1. Ray, W. H. Practical Benefits from Modelling Olefin Polymerization Reactors. In Transition Metal Catalyzed Polymerization; Quirk, R. P., Ed.; Harwood: New York, 1988. Resconi, L.; Piemontesi, F.; Franciscono, G.; Abis, L.; Fiorani, T. Olefin Polymerization at Bis(pentamethylcyclopentadienyl)Zirconium and -Hafnium Centers: Chain-Transfer Mechanisms. J. Am. Chem. Soc. 1992, 114, 1025. Richards, J. R.; Congalidis, J. P.; Gilbert, R. G. Mathematical Modeling of Emulsion Copolymerization Reactors. J. Appl. Polym. Sci. 1989, 37, 2727. Rincon-Rubio, L. M.; Wilen, C. E.; Lindfors, L. E. A Kinetic Model for the Polymerization of Propylene over a Ziegler-Natta Catalyst. Eur. Polym. J. 1990, 26, 171. Rossignoli, P. J.; Duever, T. A. The Estimation of Copolymer Reactivity Ratios: A Review and Case Studies Using the Errorin-Variables Model and Nonlinear Least Squares. Polym. React. Eng. 1995, 3, 361. Russell, G. T. On Exact and Approximate Methods of Calculating an Overall Termination Rate Coefficient from Chain Length Dependent Termination Rate Coefficients. Macromol. Theory Simul. 1994, 3, 439. Russell, G. T.; Napper, D. H.; Gilbert, R. G. Termination in FreeRadical Polymerizing Systems at High Conversion. Macromolecules 1988a, 21, 2133. Russell, G. T.; Napper, D. H.; Gilbert, R. G. Initiator Efficiencies in High-Conversion Bulk Polymerization. Macromolecules 1988b, 21, 2141. Russell, G. T.; Gilbert, R. G.; Napper, D. H. Chain-LengthDependent Termination Rate Processes in Free-Radical Polymerizations. I. Theory. Macromolecules 1992, 25, 2459. Saenz de Buruaga, I.; Arotena, M.; Armitage, P. D.; Gugliotta, L. M.; Leiza, J. R.; Asua, J. M. On-Line Calorimetric Control of Emulsion Polymerization Reactors. Chem. Eng. Sci. 1996, 51, 2781. Sarkar, P.; Gupta, S. K. Modelling of Propylene Polymerization in an Isothermal Slurry Reactor. Polymer 1991, 32, 2842 Sarkar, S.; Adhikari, M. S.; Banerjee, M.; Konar, R. S. Persulfate Initiated Aqueous Polymerization of Acrylonitrile at 50 C in an Inert Atmosphere of Nitrogen Gas. J. Appl. Polym. Sci. 1988, 36, 1865. Sarkar, S.; Adhikari, M. S.; Banerjee, M.; Konar, R. S. Mechanism of Persulfate Decomposition in Aqueous Solution at 50 C in the Presence of Vinyl Acetate and Nitrogen. J. Appl. Polym. Sci. 1990, 39, 1061. Sau, M.; Gupta, S. K. Modelling of a Semibatch Polypropylene Slurry Reactor. Polymer 1993, 34, 4417. Schmeal, W. R.; Street, J. R. Polymerization in Expanding Catalyst Particles. AIChE J. 1971, 17, 1189. Schmeal, W. R.; Street, J. R. Polymerization Catalyst Particles: Calculation of Molecular Weight Distribution. J. Polym. Sci., Polym. Phys. Ed. 1972, 10, 2173. Schoonbrood, H. A. S.; Thijssen, H. A.; Brouns, H. M. G.; Peters, M.; German, A. L. Semicontinuous Emulsion Copolymerization to Obtain Styrene-Methyl Acrylate Copolymers with Predetermined Chemical Composition Distributions. J. Appl. Polym. Sci. 1993, 49, 2029. Schoonbrood, H. A. S.; Van Den Boom, M. A. T.; German, A. L.; Hutovic, J. Multimonomer Partitioning in Latex Systems with

1014 Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997


Moderately Water-Soluble Monomers. J. Polym. Sci., Part A: Polym. Chem. 1994, 32, 2311. Schoonbrood, H. A. S.; van Eijnatten, R. C. P. M.; German, A. L. Emulsion Co- and Terpolymerization of Styrene, Methyl Methacrylate and Methyl Acrylate. II. Control of Emulsion Terpolymer Microstructure with Optimal Addition Profiles. J. Polym. Sci., Part A: Polym. Chem. 1996, 34, 949. Scott, G. E.; Senogles, E. Polymerization Kinetics of n-Lauryl Acrylate. J. Macromol. Sci., Chem. 1970, A4, 1105. Scott, G. E.; Senogles, E. Polymerization Kinetics of n-Alkyl Acrylates. J. Macromol. Sci., Chem. 1974, A8, 753. Simonazzi, T.; Cecchin, G.; Mazzaullo, S. An Outlook on Progress in Polypropylene-Based Polymer Technology. Prog. Polym. Sci. 1991, 16, 303. Singh, D.; Merrill, R. P. Molecular Weight Distribution of Polyethylene Produced by Ziegler-Natta Catalysts. Macromolecules 1971, 4, 599. Skomorokhov, V. B.; Zakharov, V. A.; Bukatov, G. D.; Kirillov, V. A.; Kryukova, G. N. Mass Transfer Processes during Polymerization of Olefins on High Efficiency Solid Catalysts. Polym. Sci. USSR 1987, 29, 979. Skomorokhov, V. B.; Zakharov, V. A.; Kirillov, V. A.; Bukatov, G. D. Mass Transfer in Polymerization of Olefins on Solid Catalysts. The Double Grain Model. Polym. Sci. USSR 1989, 31, 1420. Snuparek, J.; Kaspar, K. Semi-Continuous Emulsion Copolymerization of Ethyl Acrylate and Butyl Acrylate at High Conversions. J. Appl. Polym. Sci. 1981, 26, 4081. Soares, J. B. P. Dynamic Mathematical Modelling of Polymerization of Olefins Using Heterogeneous and Homogeneous ZieglerNatta Catalysts. Ph.D. Thesis, McMaster University, Hamilton, Ontario, Canada, 1994. Soares, J. B. P.; Hamielec, A. E. Metallocene/Aluminoxane Catalysts for Olefin Polymerization. A Review. Polym. React. Eng. 1995a, 3, 131. Soares, J. B. P.; Hamielec, A. E. General Dynamic Mathematical Modelling of Heterogeneous and Homogeneous Ziegler-Natta Copolymerization with Multiple Site Types and Mass and Heat Transfer Resistances. Polym. React. Eng. 1995b, 3, 261. Soares, J. B. P.; Hamielec, A. E. Deconvolution of Chain Length Distributions of Linear Polymers Made by Multiple Site Type Catalysts. Polymer 1995c, 36, 2257. Soares, J. B. P.; Hamielec, A. E. Effect of Reactor Residence Time Distribution on the Size of Polymer Particles Made with Heterogeneous Ziegler-Natta and Supported Metallocene Catalysts. A Generic Mathematical Model. Macromol. Chem. Theory Simul. 1995d, 4, 1085. Soares, J. B. P.; Hamielec, A. E. Analyzing TREF Data by Stockmayers Bivariate Distribution. Macromol. Chem. Theory Simul. 1995e, 4, 305. Soares, J. B. P.; Hamielec, A. E. Effect of Hydrogen and of Catalyst Prepolymerization with Propylene on the Polymerization Kinetics of Ethylene with a Non-Supported Heterogeneous ZieglerNatta Catalyst. Polymer 1996a, 37, 4599. Soares, J. B. P.; Hamielec, A. E. Kinetics of Propylene Polymerization with Non-Supported Heterogeneous Ziegler-Natta CatalystsEffect of Hydrogen on Rate of Polymerization, Stereoregularity, and Molecular Weight Distribution. Polymer 1996b, 37, 4606. Soares, J. B. P.; Hamielec, A. E. Bivariate Chain Length and Long Chain Branching Distribution for Copolymerization of Olefins and Polyolefin Chains Containing Terminal Double-Bonds. Macromol. Chem. Theory Simul. 1996c, 5, 547. Soh, S. K.; Sundberg, D. C. Diffusion-Controlled Vinyl Polymerization. I. The Gel Effect. J. Polym. Sci., Polym. Chem. Ed. 1982a, 20, 1299. Soh, S. K.; Sundberg, D. C. Diffusion-Controlled Vinyl Polymerization. III. Free Volume Parameters and Diffusion-Controlled Propagation. J. Polym. Sci., Polym. Chem. Ed. 1982b, 20, 1331. Soh, S. K.; Sundberg, D. C. Diffusion-Controlled Vinyl Polymerization. IV. Comparison of Theory and Experiment. J. Polym. Sci., Polym. Chem. Ed. 1982c, 20, 1345. Song, Z.; Poehlein, G. W. Particle Formation in Emulsion Polymerization: Transient Particle Concentration. J. Macromol. Sci., Chem. 1988a, A25, 403. Song, Z.; Poehlein, G. W. Particle Formation in Emulsion Polymerization: Particle Number at Steady State. J. Macromol. Sci., Chem. 1988b, A25, 1587. Spitz, R. Molecular Weight Distribution in Ziegler-Natta Olefin Polymerization. In Recent Advances in Mechanistic and Synthetic Aspects of Polymerization; Fontanille, M., Guyot, A., Eds.; D. Reidel Publishing Co.: New York, 1987. Stickler, M. Free-Radical Polymerization Kinetics of Methyl Methacrylate at Very High Conversions. Makromol. Chem. 1983, 184, 2563. Stickler, M.; Panke, D.; Hamielec, A. E. Polymerization of Methyl Methacrylate up to High Degrees of Conversion: Experimental Investigation of the Diffusion-Controlled Polymerization. J. Polym. Sci., Polym. Chem. Ed. 1984, 22, 2243. Stockmayer, W. H. Distribution of Chain Lengths and Compositions in Copolymers. J. Chem. Phys. 1945, 13, 199. Stockmayer, W. H. Note on the Kinetics of Emulsion Polymerization. J. Polym. Sci. 1957, 24, 314. Storti, G.; Morbidelli, M.; Carra, S. Detailed Modelling of Multi component Emulsion Polymerization Systems. Polym. Mater. Sci. Eng. 1988, 58, 600. Storti, G.; Carra, S.; Morbidelli, M.; Vita, G. Kinetics of Multi` monomer Emulsion Polymerization. The Pseudo-Homopolymerization Approach. J. Appl. Polym. Sci. 1989, 37, 2443. Sundberg, D. C. A Quantitative Treatment of Particle Size Distributions in Emulsion Polymerization. J. Appl. Polym. Sci. 1979, 23, 2197. Sutton, T. L.; MacGregor, J. F. The Analysis and Design of Binary Vapour-Liquid Equilibrium Experiments. Part I: Parameter Estimation and Consistency Tests. Can. J. Chem. Eng. 1977, 55, 602. Tacx, J. C. J. F.; Linssen, H. N.; German, A. L. Effect of Molar Mass Ratio of Monomers on the Mass Distribution of Chain Lengths and Compositions in Copolymers: Extension of Stockmayer Theory. J. Polym. Sci., Part A: Polym. Chem. 1988, 26, 61. Tait, P. J. T. Monoalkene Polymerization: Ziegler-Natta and Transition Metal Catalysts. In Comprehensive Polymer Science, Allen, Sir G., Ed.; Pergamon Press: Oxford, U.K., 1989. Tait, P. J. T.; Wang, S. Studies on the Polymerization of Propylene Using High Activity Ziegler-Natta Catalyst. 1. Kinetic Models for Rate Decay in Ziegler-Natta Polymerization. Br. Polym. J. 1988, 20, 499. Tait, P. J. T.; Watkins, N. D. Monoalkene Polymerization Mechanisms. In Comprehensive Polymer Science; Allen, Sir G., Ed.; Pergamon Press: Oxford, U.K., 1989. Taylor, T. W.; Choi, K. Y.; Yuan, H.; Ray, W. H. Physicochemical Kinetics of Liquid Phase Propylene Polymerization. MMI Press Symp. Ser. 1983, 4, 191. Teymour, F.; Campbell, J. D. Analysis of the Dynamics of Gelation in Polymerization Reactors Using the Numerical Fractionation Technique. Macromolecules 1994, 27, 2460. Tidwell, P. W.; Mortimer, G. A. An Improved Method of Calculating Copolymerization Reactivity Ratios. J. Polym. Sci. A 1965, 3, 369. Tidwell, P. W.; Mortimer, G. A. Science of Determining Copolymerization Reactivity Ratios. J. Macromol. Sci., Rev. 1970, C4, 281. Tobita, H. Cross-Linking Kinetics in Emulsion Copolymerization. Macromolecules 1992, 25, 2671. Tobita, H. Molecular Weight Distribution in Free-Radical CrossLinking Copolymerization. Macromolecules 1993a, 26, 836. Tobita, H. Kinetics of Network Formation in Free-Radical CrossLinking Copolymerization. Macromolecules 1993b, 26, 5427. Tobita, H. A Simulation Model for Long-Chain Branching in Vinyl Acetate Polymerization: 1. Batch Polymerization. J. Polym. Sci., Part B: Polym. Phys. 1994a, 32, 901. Tobita, H. A Simulation Model for Long-Chain Branching in Vinyl Acetate Polymerization: 2. Continuous Polymerization in a Stirred Tank Reactor. J. Polym. Sci., Part B: Polym. Phys. 1994b, 32, 911. Tobita, H. Monte Carlo Simulation of Emulsion PolymerizationsLinear, Branched, and Crosslinked Polymers. Acta Polym. 1995a, 46, 185. Tobita, H. Molecular Weight Distribution in Random Crosslinking of Polymer Chains. J. Polym. Sci., Polym. Phys. 1995b, 33, 1191. Tobita, H.; Hamielec, A. E. Modeling of Network Formation in Free Radical Polymerization. Macromolecules 1989, 22, 3098. Tobita, H.; Hamielec, A. E. Kinetics of Free-Radical Copolymerization: The Pseudo-Kinetic Rate Constant Method. Polymer 1991, 32, 2641. Tobita, H.; Hamielec, A. E. Control of Network Structure in FreeRadical Crosslinking Copolymerization. Polymer 1992, 33, 3647. Ugelstad, J.; Hansen, F. K. Kinetics and Mechanism of Emulsion Polymerization. Rubber Chem. Technol. 1976, 49, 536.

Ind. Eng. Chem. Res., Vol. 36, No. 4, 1997 1015


Ugelstad, J.; Mrk, P. C.; Aasen, J. O. Kinetics of Emulsion Polymerization. J. Polym. Sci., Polym. Chem. Ed. 1967, 5, 2281. Ugelstad, J.; El-Aasser, M.; Vanderhoff, J. Emulsion Polymerizations: Initiation of Polymerization in Monomer Droplets. J. Polym. Sci., Polym. Lett. 1973, 11, 503. Ugelstad, J.; Hansen, F. K.; Lange, S. Emulsion Polymerization of Styrene with Sodium Hexadecyl Sulphate/Hexadecanol Mixtures as Emulsifiers. Initiation in Monomer Droplets. Makromol. Chem. 1974, 175, 507. Urquiola, B.; Arzamendi, F.; Leiza, J. R.; Zamora, A.; Asua, J. M.; Delgado, J.; El-Aasser, M. S.; Vanderhoff, J. W. Semicontinuous Seeded Emulsion Copolymerization of Vinyl Acetate and Methyl Acrylate. J. Polym. Sci., Part A: Polym. Chem. 1991, 29, 169. Urretabizkaia, A.; Sudol, E. D.; El-Aasser, M. S.; Asua, J. M. Calorimetric Monitoring of Emulsion Copolymerization Reactions. J. Polym. Sci., Part A: Polym. Chem. 1993, 31, 2907. Urretabizkaia, A.; Arzamendi, G.; Asua, J. M. Modeling Semicontinuous Emulsion Terpolymerization. Chem. Eng. Sci. 1992, 47, 2579. Urretabizkaia, A.; Asua, J. M. High Solids Content Emulsion Terpolymerization of Vinyl Acetate, Methyl Methacrylate and Butyl Acrylate. I. Kinetics. J. Polym. Sci., Part A: Polym. Chem. 1994, 32, 1761. Urretabizkaia, A.; Arzamendi, G.; Unzue, M. J.; Asua, J. M. High Solids Content Emulsion Terpolymerization of Vinyl Acetate, Methyl Methacrylate and Butyl Acrylate. II. Open Loop Composition Control. J. Polym. Sci., Part A: Polym. Chem. 1994a, 32, 1779. Urretabizkaia, A.; Leiza, J. R.; Asua, J. M. On-line Terpolymer Composition Control in Semicontinuous Emulsion Polymerization. AIChE J. 1994b, 40, 1850. Usami, T.; Gotoh, Y.; Takayama, S. Generation Mechanisms of Short-Chain Branching Distribution in Linear Low-Density Polyethylenes. Macromolecules 1986, 19, 2722. Van der Meer, R.; Linssen, H. N.; German, A. L. Improved Methods of Estimating Monomer Reactivity Ratios in Copolymerization by Considering Experimental Errors in Both Variables. J. Polym. Sci., Polym. Chem. Ed. 1978, 16, 2915. van Doremaele, G. H. J.; Schoonbrood, H. A. S.; Kurja, J.; German, A. L. Copolymer Composition Control by Means of Semicontinuous Emulsion Copolymerization. J. Appl. Polym. Sci. 1992, 45, 957. Vela-Estrada, J. M.; Hamielec, A. E. Modelling of Ethylene Polymerization with (Cp)2ZrCl2/MAO Catalyst. Polymer 1994, 35, 808. Vickroy, V. V.; Schneider, H.; Abbott, R. F. The Separation of SEC Curves of HDPE into Flory Distributions. J. Appl. Polym. Sci. 1993, 50, 551. Vivaldo-Lima, E.; Hamielec, A. E.; Wood, P. E. Auto-Acceleration Effect in Free Radical Polymerization. A Comparison of the CCS and MH Models. Polym. React. Eng. 1994a, 2, 17. Vivaldo-Lima, E.; Hamielec, A. E.; Wood, P. E. Batch Reactor Modelling of the Free Radical Copolymerization Kinetics of Styrene/Divinylbenzene up to High Conversions. Polym. React. Eng. 1994b, 2, 87. Welborn, H. C., Jr. Supported polymerization catalyst. U.S. Patent 4,701,432, 1987. Welborn, H. C., Jr. New supported polymerization catalyst. U.S. Patent 5,077,255, 1991. Welborn, H. C., Jr. Supported polymerization catalyst. U.S. Patent 5,124,418, 1992. Welborn, H. C., Jr. Polymerization process using a new supported polymerization catalyst. U.S. Patent 5,183,867, 1993. Whiteley, K. S.; Heggs, T. G.; Koch, H.; Mawer, R. L.; Immel, W. Polyolefins. In Ullmanns Encyclopedia of Industrial Chemistry; Elvers, B., Hawkins, S., Schultz, G., Eds.; VCH Publishers Inc.: Weinheim, Germany, 1992; Vol. A21, p 487. Wilchinsky, Z. W.; Looney, R. W.; Tornquist, G. M. Dependence of Polymerization Activity on Particle and Crystallite Dimensions in Ball Milled TiCl3 and TiCl3/0.33AlCl3 Catalyst Components. J. Catal. 1973, 28, 351. Xie, T.; Hamielec, A. E. Modelling Free-Radical Copolymerization KineticssEvaluation of the Pseudo-Kinetic Rate Constant Method, 1. Molecular Weight Calculation for Linear Copolymers. Makromol. Chem., Theory Simul. 1993a, 2, 421. Xie, T.; Hamielec, A. E. Modelling Free-Radical Copolymerization KineticssEvaluation of the Pseudo-Kinetic Rate Constant Method, 2. Molecular Weight Calculations for Copolymers with Long Chain Branching. Makromol. Chem., Theory Simul. 1993b, 2, 455. Xie, T.; Hamielec, A. E. Modelling Free-Radical Copolymerization Kinetics, 3. Molecular Weight Calculations for Copolymers with Crosslinking. Makromol. Chem., Theory Simul. 1993c, 2, 777. Xie, T. Y.; Hamielec, A. E.; Wood, P. E.; Woods, D. R. Experimental Investigation of Vinyl Chloride Polymerization at High Conversion: Semi-Batch Reactor Modelling. Polymer 1991, 32, 2087. Xie, T.; McAuley, K. B.; Hsu, J. C. C.; Bacon, D. W. Gas Phase Ethylene Polymerization: Production Processes, Polymer Properties, and Reactor Modelling. Ind. Eng. Chem. Res. 1994, 33, 449. Yamada, B.; Itahashi, M.; Otsu, T. Estimation of Monomer Reactivity Ratios by Nonlinear Least-Squares Procedure with Consideration of the Weight of Experimental Data. J. Polym. Sci. 1978, 16, 1719. Yaraskavitch, I. M.; Brash, J. L.; Hamielec, A. E. An Investigation of the Kinetics of Copolymerization of p-Methylstyrene/Acrylonitrile to High Conversion: Modelling Diffusion-Controlled Termination and Propagation by Free-Volume Theory. Polymer 1987, 28, 489. Zhu, S.; Hamielec, A. E. Chain-Length-Dependent Termination for Free Radical Polymerization. Macromolecules 1989, 22, 3093. Zhu, S.; Tian, Y.; Hamielec, A. E.; Eaton, D. R. Radical Trapping and Termination in Free-Radical Polymerization of MMA. Macromolecules 1990a, 23, 1144. Zhu, S.; Tian, Y.; Hamielec, A. E.; Eaton, D. R. Radical Concentrations in Free Radical Copolymerization of MMA/EGDMA. Polymer 1990b, 31, 154. Zucchini, U.; Cecchin, G. Control of Molecular Weight Distribution in Polyolefins Synthesized with Ziegler-Natta Catalytic Systems. Adv. Polym. Sci. 1983, 51, 101.

Received for review August 1, 1996 Revised manuscript received January 10, 1997 Accepted January 10, 1997X IE960481O

X Abstract published in Advance ACS Abstracts, March 1, 1997.

Vous aimerez peut-être aussi