Sie sind auf Seite 1von 16

Vol.

2 CRITICAL PHASE POLYMERIZATIONS 111

CRITICAL PHASE POLYMERIZATIONS


Introduction

A supercritical uid (SCF) can be dened as a substance or mixture which has


both a temperature and pressure that exceed its critical temperature (T c ) and
critical pressure (Pc ), and a density at or above its critical density (1) (Fig. 1).
Near this critical point, the density, transport properties (such as diffusivity and
viscosity), and other physical properties (such as dielectric constant and solvent
strength) can be varied in a continuum from gas-like to liquid-like, with relatively
small changes in temperature or pressure. The tunability of SCFs makes them
very attractive as solvents for polymerization reactions.
The rst example of polymerization in an SCF and the most commercially
signicant till this time is the high pressure synthesis of low density polyethylene
(PE) (2). In this process, ethylene from a recycle stream and a makeup source are
compressed to 250 MPa (2500 bar) in a tubular reactor at 250 C. The free-radical
reaction is initiated with either peroxide or traces of oxygen. The polymer pro-
duced has an average molecular weight higher than 100,000 g/mol that is soluble
in the ethylene monomer at the process conditions. Polyethylene is precipitated
and collected in a separator downstream of the reactor section by reducing the
pressure of the solution to 35 MPa (350 bar), thus changing the solvent quality of
ethylene.
In an effort to reduce emission of volatile organic compounds, to completely
phase out the use of chlorouorocarbons (CFCs), and to reduce the generation of
aqueous waste streams generated in the polymer industry, scientists have turned
to the use of supercritical carbon dioxide (scCO2 ). In the 1990s, CO2 has proven
to be a suitable alternative solvent for critical phase polymerization reactions
(3). It has an easily accessible critical point with a T c of 31.1 C and a Pc of 7.38 MPa

Encyclopedia of Polymer Science and Technology. Copyright John Wiley & Sons, Inc. All rights reserved.
112 CRITICAL PHASE POLYMERIZATIONS Vol. 2

Supercritical

Fluid

Region
7380
liquid
Pressure, kPa solid

530

vapor

57.0 31.1
Temperature, C

Fig. 1. Supercritical uid phase diagram. To convert kPa to bar, divide by 100.

1.25
43C
23C
3C
1.00 17C
37C
57C
Density, g/mL

0.75 77C
97C

0.50

0.25

0.00
10 20 30
Pressure, MPa

Fig. 2. CO2 density prole. To convert MPa to psi, multiply by 145.

(73.8 bar) (4). Within this region, scCO2 exists as a low viscosity medium with a
tunable density (Fig. 2). It is nontoxic, nonammable, relatively inexpensive, nat-
urally abundant, and generally regarded as safe. Because CO2 is an ambient gas,
it can be separated from the polymer by depressurizing the reaction vessel, result-
ing in a dry polymer product. CO2 is also inert toward free radicals and cations,
which lessens chain-transfer reactions. These aspects make CO2 very attractive
Vol. 2 CRITICAL PHASE POLYMERIZATIONS 113

for use in synthesis of industrially important polymers. Numerous examples exist


of both chain-growth and step-growth polymerizations using CO2 as a reaction
medium. This article focuses on the utilization of liquid and scCO2 as a reaction
medium for polymer synthesis.
Knowledge of two important issues is important when CO2 is used as a
polymerization solvent: solubility and plasticization. CO2 is a nonpolar solvent
and should generally dissolve anything that dissolves in hexane. This concept is
valid for many low molar mass compounds that have appreciable vapor pressures;
however, it fails in the case of polymers that have negligible vapor pressures (5).
In fact, the only polymers that show good solubility in CO2 under mild conditions
[<100 C, <35 MPa (<350 bar)] are amorphous uoropolymers, siloxanes (6,7),
and polyether carbonates (8). CO2 has a very low polarizability per unit volume
(27.6 10 25 cm3 ) (9) and a low dielectric constant (1.011.67) (10), making
its solvating power similar to that of a uorocarbon oil. Because of structural
symmetry, CO2 lacks a dipole moment; however, it does have a strong quadrapole
moment that contributes to its solubility parameter. This quadrapole moment
along with the Lewis acidity of CO2 allows it to participate in specic solute
solvent interactions that allow certain polymers to have enhanced solubilities
in CO2 . Experiments conducted using Fourier transform infrared spectroscopy
reveals that CO2 is capable of having Lewis acidbase type interactions with
polymers that have electron-donating functional groups such as carbonyls (11).
Amorphous uoropolymers have an especially high solubility that is attributed to
specic interactions such as the formation of weak CO2 uoropolymer complexes
or favored clustering of the CO2 near the more polar C F bonds of the polymer.
High resolution 1 H and 19 F nmr studies conrm that such interactions do exist,
thus causing magnetic shielding of the uorine nuclei (12,13).
The unique solubility characteristics of CO2 limit the types of polymerization
techniques that can be successfully employed. While amorphous uoropolymers
may be synthesized homogeneously in CO2 , most other polymers are insoluble
in CO2 and must be made via heterogeneous processes such as precipitation,
dispersion, and emulsion techniques or via mass polymerizations swollen by CO2 .
The plasticization of polymers by CO2 , which causes a decrease in the glass-
transition temperature (T g ) of the polymer, is another important feature that must
be taken into account. This plasticization facilitates the occurrence of important
effects that are essential to polymer synthesis. Plasticization of the polymer al-
lows enhanced diffusion of monomer and initiator into the polymer phase which
often results in increased polymerization rates in heterogeneous polymerizations,
removal of residual monomer, solvent, or catalyst from the polymer, and the forma-
tion of blends by polymerization within a CO2 -swollen host polymer. An important
emerging issue is the role that CO2 plays in manipulating the loci of reactants
such as monomer/comonomer/initiator partition coefcients. Each of these topics
is discussed in this chapter.

Chain-Growth Polymerizations in CO2

Homogeneous Polymerizations. Fluoropolymers are used in many


technologically demanding applications because of their balance of high
114 CRITICAL PHASE POLYMERIZATIONS Vol. 2

performance properties. Several signicant impediments to the synthesis of com-


mercially important uoropolymers exist including their general insolubility in
most solvents, the undesirable effect of chain transfer of uoroolen radicals to
most hydrogen-containing solvents (hence the extensive use of CFC solvents),
and their high hydrophobicity (which necessitates the use of specially uorinated
surfactants when synthesized in water). Preparing these amorphous low melt-
ing materials in scCO2 can circumvent the environmental concerns for CFCs and
can avoid the ensemble of problems associated with their synthesis in water. The
feasibility of utilizing liquid and scCO2 as a solvent for homogeneous solution
polymerization of high molar mass amorphous uoropolymers by using thermal
free-radical initiation has been demonstrated (1416). Several uorinated acry-
lates have been polymerized using this methodology to give high yields of polymer.
For example, poly(1,1-dihydroperuoroctyl acrylate) (PFOA) (50 w/v) was poly-
merized using 2,2 -azobis(isobutyronitrile) (AIBN) as the initiator at 59.4 C and
20.7 MPa (207 bar) CO2 for 48 h (eq. (1)).

(1)

The results from a uv spectroscopy study revealed that the decomposition


rate of AIBN in CO2 is 2.5 times slower than those observed in more viscous
solvents (15). This was attributed to the difference in dielectric constants of the
solvents. In contrast, higher initiator efciencies were observed in CO2 , relative
to other solvents, as a result of a decreased solvent cage effect in the supercriti-
cal phase due to the inherently low solvent viscosity. Other uorinated acrylates,
methacrylates, and peruoroalkyl derivatized styrenes have been homogeneously
polymerized or copolymerized in CO2 (1719). The free-radical propagation ki-
netics of styrene, methyl methacrylate, and butyl methacrylate in CO2 has been
evaluated using a pulsed laser polymerization (PLP) (2022). The propagation rate
was found to be slightly lower than that in the bulk, which indicates that CO2 does
not interfere with the chain-growth process. Homogeneous free-radical polymer-
izations have also been used to synthesize telomers of 1,1-diuoroethylene (23).
The peruoroalkyl iodides prepared in this manner had both controlled molecular
weights and relatively narrow molecular weight distributions.
Precipitation Polymerizations. A variety of polymers have been syn-
thesized in CO2 by free-radical precipitation polymerization. In the earliest
studies, the free-radical polymerization of ethylene in CO2 using either gamma
radiation or AIBN initiation was investigated (2426). Other patents (27,28) dis-
closed the polymerization of several vinyl hydrocarbon monomers in liquid and
scCO2 . Homopolymers of vinyl chloride, acrylonitrile, acrylic acid, vinyl acetate,
2-hydroxyethylacrylate, and some N-vinylcarboxamides as well as PS/PMMA
and PVC/PVAc random copolymers have been made. The polymerization re-
actions resulted in polymers with gravimetric yields ranging from 15 to 97%
and viscosity average molecular weights (M v ) ranging from 1.2 104 to 1.6
106 g/mol depending on the monomer and the reaction conditions used. More re-
cently, the free-radical polymerization of high molecular weight poly(acrylic acid)
Vol. 2 CRITICAL PHASE POLYMERIZATIONS 115

(M n = 1.5105 g/mol) has been achieved in scCO2 despite the fact that the polymer
precipitates from solution at molecular weights greater than 1000 g/mol (29,30).
The molecular weight of the polymer could be controlled by either adding ethyl
mercaptan as a chain-transfer agent or by manipulating the temperature and
pressure parameters for the reaction. High performance uoropolymers have also
been synthesized in CO2 by free-radical precipitation polymerization. Homo- and
copolymers of tetrauoroethylene (TFE) have generated particular interest, since
it was discovered that TFE can be handled more safely as a mixture with CO2 (31).
In 1995, homopolymerizations (32) of TFE and copolymerizations (33,34) of TFE
with hexauoropropylene (HFP) or peruoropropyl vinyl ether (PPVE) yielding
high molecular weight polymers (>106 g/mol) in good yield were reported. Copoly-
merization of TFE and peruoroalkyl vinyl ether monomers in CO2 occurs with vir-
tually no -scission, implying that propagation of terminal vinyl ether-containing
radicals is more favorable in CO2 than in traditional solvents. A system for the con-
tinuous polymerization of various monomers in scCO2 has been developed (35).
Continuous precipitation polymerization of vinylidene uoride and acrylic acid
can occur in an intensely agitated continuous stirred tank reactor (CSTR). Rates
of polymerization of vinylidene uoride in the CSTR are signicantly higher than
the average rates of batch polymerization, under similar conditions. CO2 has also
been employed as a solvent in precipitation polymerizations of cross-linked poly-
mers, using high levels of divinylbenzene cross-linker (36). The AIBN-initiated
reactions were performed at 65 C for 12 h in 31 MPa (310 bar) CO2 and yielded
spherical particles without the use of surfactants. Multifunctional monomers such
as trimethylolpropane trimethacrylate and ethylene glycol dimethacrylate were
also investigated (37). These reactions produced well-dened polymer monoliths
with pore diameters ranging from 20 to 8000 nm (see the Porous Polymer Syn-
thesis section). A clear advantage to the use of CO2 in each of these precipitation
reactions is that the products are easily isolated from the depressurized reaction
vessels as dry, solvent-free materials.
Dispersion and Emulsion Polymerizations. The relative insolubility
of many industrially important polymers in scCO2 necessitates the use of het-
erogeneous polymerization techniques conducted with added stabilizers. Conven-
tional heterogeneous dispersion polymerizations of unsaturated monomers are
performed in either aqueous or organic dispersing media, with the addition of
interfacially active agents to stabilize the colloidal dispersion that forms (38,39).
Successful stabilization of the polymer colloid during polymerization results in
the formation of high molar mass polymers with high rates of polymerization. A
similar analogy applies to dispersion polymerization in CO2 where stabilization
can be accomplished with the aid of nonionic homopolymers, block copolymers,
and reactive macromonomer surfactants (40) (Fig. 3). Typically, a free-radical
dispersion polymerization in scCO2 starts as a one-phase, homogeneous system
such that both the monomer and the polymerization initiator are soluble in the
polymerization medium. As the reaction progresses, the resulting growing poly-
mer phase separates into primary particles that are stabilized by amphiphilic
materials so as to prevent particle occulation and aggregation. Polymer colloids
produced by dispersion polymerizations in CO2 are usually stabilized by a steric
mechanism as compared with an electrostatic mechanism that is common to col-
loidal stabilization in aqueous environments because of the low dielectric con-
stant of CO2 . Steric stabilization of a colloidal dispersion is usually imparted by
116 CRITICAL PHASE POLYMERIZATIONS Vol. 2

Fig. 3. Surfactants for CO2 .


Vol. 2 CRITICAL PHASE POLYMERIZATIONS 117

amphiphilic macromolecules that become adsorbed onto the surface of the dis-
persed phase. These macromolecules contain an anchoring segment that attaches
to the particle usually by physical adsorption and stabilizing moieties that are sol-
uble in the continuous phase. The stabilizer prevents aggregation of particles by
coating the surface of each particle and imparting long-range repulsions between
them.
A substantial number of studies on the dispersion polymerization of lipophilic
monomers in CO2 have focused on methyl methacrylate (MMA). The rst success
in this area exploited the amphiphilic nature of the homopolymer PFOA as a
stabilizer (40) (eq. (2)).

(2)

MMA was polymerized heterogeneously to very high conversions (>90%)


and high degrees of polymerization (>3000 g/mol) in scCO2 . These polymeriza-
tions were conducted at 65 C and 20.7 MPa (207 bar) CO2 with AIBN or uo-
rinated derivatives of AIBN as initiators and PFOA (M n = 1.1 104 or 2.0
105 g/mol) as the surfactant. The reactions resulted in kinetically stable disper-
sions of micrometer-sized particles with a narrow size distribution. In contrast, re-
actions that were performed without PFOA as a stabilizer resulted in precipitated
polymer with an unstructured, nonspherical morphology. Several other monomers
were also polymerized by dispersion polymerization methods in CO2 (Table 1).
Table 1. Summary of Dispersion and Emulsion Polymerizations in CO2
Monomer(s) Stabilizers Reference
Acrylamide Amide end-capped PFPE 41
2,6-Dimethylphenylene PFOA, PFOA-based random 42
oxide copolymers, PS-b-PFOA
Divinylbenzene and Fluorinated methacrylatePMMA 37
ethylvinylbenzene block copolymer
Methyl methacrylate PFOA 40,43,44
PDMS macromonomer (45,46)(47)
Fluorinated graft or block copolymers 48
Styrene PFOA, PS-b-PFOA 49
PDMS, PS-b-PDMS 50
FVE-b-MVE 51
Carboxylic acid terminated PFPE 52
Vinyl acetate PFOA, PDMS, PDMS macromonomer, 53
PVAc-b-PDMS and PVAc-b-PFOA
Vinyl acetate and ethylene PDMS, PVAc-b-PFOA 53
1-Vinyl-2-pryrrolidinone PFOA 54
Acrylonitrile PFOA, PS-b-PFOA 55
118 CRITICAL PHASE POLYMERIZATIONS Vol. 2

Typical emulsion polymerizations utilize oil-soluble monomers dispersed in


an aqueous media with a water-soluble initiator, whereas inverse emulsions
employ water-soluble monomers dispersed in an organic medium containing an
oil-soluble initiator. The insoluble polymer particles that result from these re-
actions are stabilized as colloids in solution by repulsive forces imparted by a
small molecule ionic surfactant and/or an amphiphilic macromolecular stabilizer
(56,57). These reactions produce high molecular weight, spherical polymer par-
ticles with sizes typically smaller than 1 m. Since most common monomers
are highly soluble in CO2 , there are very few examples of CO2 -based emulsion
polymerizations. However, acrylamide is a rare example of a vinyl monomer that
has a low solubility in CO2 at moderate temperatures and pressures. The AIBN-
initiated water-in-oil or inverse emulsion polymerization of acrylamide has been
attempted at 65 C in 35.2 MPa (352 bar) CO2 (41,58) (eq. (3)).

(3)

Reactions performed with an amide end-capped peruoropolyether (PFPE)


appeared to form a milky-white latex, but produced polymers with conversions
and molecular weights similar to those polymers precipitated in reactions with-
out a stabilizer. More inverse emulsion studies are expected with the utiliza-
tion of surfactants that have the ability to form water in CO2 microemulsions.
Several researchers have demonstrated that various amphiphilic surfactants
like the graft copolymer PFOA-g-PEO (59), an ammonium carboxylate PFPE
[(OCF2 -CF(CF3 ))n (OCF2 )m ]OCF2 COO NH4 + (60), and the hybrid hydrocarbon
uorocarbon C7 F15 CH(OSO3 Na+ )C7 H15 (61) are able to stabilize water-in-CO2
microemulsions. In principle, these same surfactants could be used to form mi-
croemulsions of water and water-soluble monomers in CO2 for inverse emulsion
polymerizations.
In Situ Synthesis of Polymer Blends. The ability of scCO2 to plasticize
most polymers enables the synthesis of polymer blends that are not easily accessi-
ble by other methods. Supercritical CO2 can be used to swell a CO2 -insoluble poly-
mer matrix and subsequently infuse CO2 -soluble monomers and initiators into the
matrix for polymerization. In one system, poly(chlorotriuorethylene) (PCTFE),
poly(4-methyl-1-pentene) (PMP), high density polyethylene (HDPE), nylon-6,6-
poly(oxymethylene), and bisphenol A polycarbonate (PC) were each swollen with
CO2 and infused with styrene monomer and either AIBN or tert-butyl perbenzoate
initiators (62,63). Polymerization was thermally initiated within the polymer ma-
trix either in the presence of CO2 or N2 to generate a polymer blend. Mass uptakes
of up to 118% based on the original mass of the polymer were observed. In the
case of PCTFE, extraction of polystyrene from the matrix and subsequent gpc anal-
ysis showed high molecular weight polymer (M n > 1 105 g/mol) in the blend.
Transmission electron microscopy and energy dispersive x-ray analysis were used
to demonstrate that the polystyrene exists as discrete phase-segregated regions
Vol. 2 CRITICAL PHASE POLYMERIZATIONS 119

throughout the matrix polymer. Thermal analysis results revealed that radical
grafting reactions are not signicant. Other materials including PS/PE compos-
ites (64) and FEP/PS blends (65) can be synthesized by related routes. In a simi-
lar study this strategy has been used as a route to polymer surface modication.
Styrene (with and without various cross-linkers) was infused into three polymer
substrates (PTFE, PCTFE, and FEP), polymerized, and subsequently sulfonated
in order to provide surface modied semi-interpenetrating networks (66). An in-
crease in wettability of the modied uoropolymer matrices was shown using
water contact angle measurements.
Cationic Polymerizations. Homogeneous and heterogeneous cationic
polymerizations of various monomers in liquid and scCO2 have been performed
(Table 2). Several different catalyst systems were employed in these polymeriza-
tions, all of which proved to be effective. The early experiments conducted using
monomers such as isobutylene (67) and formaldehyde (7072) normally resulted
in a low yield of low molecular weight products. Nonetheless, proof of concept
for the use of CO2 as an effective solvent for cationic chain-growth reactions was
achieved. Other polymers based on different substituted alkenes (73), vinyl ethers
(74,75) (eq. (4)), and oxetanes (75) (eq. (5)) have been synthesized.

Table 2. Summary of Cationic Polymerizations in CO2 a


Monomer (s) Catalyst System Reference
Isobutylene AlBr3 /ethyl bromide (cosolvent) 67
TiCl4 /isopropyl chloride 67
2-(2,4,4-trimethylpentyl) chloride 68
(TMPSCl)/TiCl4 /methyl chloride
Isobutylene/styrene TMPSCl/TiCl4 /methyl chloride 69
Ethyl vinyl ether SnCl4 or BF3 -O(C2 H5 )2 28
Formaldehyde Acetic acid or triuoroacetic acid 70,7172
3-Methyl-1-butene Water/AlCl3 /methyl chloride 73
4-Methyl-1-pentene Water/AlCl3 /methyl chloride 73
Isobutyl vinyl ether (IBVE) Acetic acid adduct of IBVE 74,75
C2 H5 AlCl2
C2 H5 OOCCH3 (Lewis base deactivator)
2-(N-propyl-N-peruorooctyl- Acetic acid adduct of FVE 74,75
sulfonamide ethyl vinyl ether
(FVE)
C2 H5 AlCl2 /C2 H5 OOCCH3
Bis(ethoxymethyl)oxetane BH3 -THF/NaOHb 76
3-Methyl-3 [1,1-dihydrohepta- BH3 -THF/NaOH 76
uorobutoxy)methyl]-oxetane
FVE/methyl vinyl ether Acetic acid adduct of FVE C2 H5 AlCl2 / 76
C2 H5 OOCCH3
Styrene TiCl4 c 77
a reactionsare homogeneous unless noted.
b heterogeneous precipitation reaction.
c heterogeneous dispersion reaction.
120 CRITICAL PHASE POLYMERIZATIONS Vol. 2

(4)

(5)

Fast reaction kinetics generated by the high reactivity of carbocations is a


major benet in cationic polymerizations. However, this same reactivity leads to
unwanted side reactions such as chain transfer to monomer and early chain ter-
mination by abstraction of acidic hydrogen atoms beta to the carbocation. CO2 has
several features that make it a suitable solvent for cationic polymerization reac-
tions. It is relatively inert toward the carbocationic propagating species and is not
incorporated into the polymer backbone (75). CO2 as a solvent mimics the typical
nonpolar solvents that are conventionally used by affecting the equilibrium be-
tween contact pairs and solvent-separated pairs. This in turn affects the activation
energy for chain transfer and termination reactions. To minimize the limitations of
cationic polymerizations, lower reaction temperatures (typically 70 C to 30 C)
as well as controlled polymerization techniques are used. Living cationic polymer-
ization methods that involve the stabilization of the cationic propagating species
with a nucleophilic counterion or an added Lewis base have been developed and
allow well-dened polymer and block copolymers to be prepared (75).
Transition-Metal-Catalyzed Polymerizations. Metal-catalyzed poly-
merizations (76) have been performed in scCO2 (77,78). The ring-opening metathe-
sis polymerization (ROMP) of norbornene has been performed with CO2 and
CO2 /methanol mixtures using a Ru(H2 O)6 (O3 SC6 H4 CH3 -p)2 catalyst as the ini-
tiator (78) (eq. (6)).

(6)

These reactions were carried out at 65 C and 634.5 MPa (60345 bar)
CO2 and yielded polymers with conversions, molecular weights, PDIs, and poly-
mer microstructures similar to those obtained in other solvent systems. Addi-
tion of methanol as a cosolvent to the polymerization medium accomplished sev-
eral things: it enabled the ruthenium catalyst that was insoluble in neat CO2
to be solubilized in the CO2 /monomer mixture; it allowed for increased rates of
Vol. 2 CRITICAL PHASE POLYMERIZATIONS 121

polymerization yielding similar conversions in a shorter period of time compared


to polymerizations done in pure CO2 ; and it strongly affected the cistrans ratio
of the polymer microstructure by favoring the trans carbene propagating species
(76,79). Both a ruthenium-based Grubbs catalyst (80) and a molybdenum-based
Shrock catalyst (81) have shown even greater activity in the ROMP of norbornene,
producing high molecular weight polymer (105 106 g/mol) in good yields (up to
97%) at milder reaction conditions [2545 C and 11 MPa (110 bar) CO2 ]. The
ruthenium catalyst was also used to polymerize cis-cyclooctene in 50% yield and
a molecular weight of 105 g/mol.
An insoluble rhodium catalyst [(nbd)Rh(acac)] has been used to synthesize
polyphenylacetylene in the presence of liquid and scCO2 (82). The catalyst is sol-
ubilized in the reaction mixture by addition of a peruoroalkyl-substituted triph-
enylphosphine ligand. Polyphenylactylene formed in good yields (6075%) had a
molecular weight of about 4 104 g/mol for THF soluble fractions, and molecular
weight distributions in the range of 810.
A number of metal-catalyzed polymerizations have utilized CO2 as both a
solvent and as a reagent in the reactions. Precipitation copolymerization of either
propylene oxide (83) or cyclohexene oxide (84) with CO2 in scCO2 has been cat-
alyzed using heterogeneous zinc catalysts. Copolymerizations of CO2 and propy-
lene oxide formed PCs with a molecular weight of about 104 g/mol and incor-
poration of CO2 at greater than 90% (eq. (7)). A small percentage of propylene
carbonate by-product was also observed.

(7)

The cyclohexene oxide reactions yielded polymers with M w = 3.8 105 g/mol
and 91% CO2 incorporation. Catalyst efciencies for these two systems were low
to average (3400 g of polymer per gram of Zn), possibly because of the insolubility
of the zinc complexes in the CO2 /monomer mixture (85). Molecular weight distri-
butions were also very broad (PDI = 2.427). A partially uorinated porphyrin
complex was found to be soluble in CO2 , and CO2 /monomer mixtures catalyzed
the polymerization of cyclohexene oxide and CO2 (in the presence of a base) form-
ing polymer in good yields with high degrees of carbonate linkages (9097%) (86).
This system demonstrated substantially higher catalyst efciencies (up to 3900 g
of polymer per gram of Cr) than other systems and very narrow molecular weight
distributions (PDI < 1.4).
Thermal Ring-Opening Polymerization. The thermal ring-opening
polymerization of a silicone-bridged [1]-ferroceneophane in scCO2 was performed
at 20.7 MPa (207 bar) and 75130 C (87) (eq. (8)). The polymer molecular weights
were of the same order of magnitude as the polymers prepared by the solvent-free
methods, but were uniformly lower (M w = 2.87 105 vs 5.2 105 g/mol).
122 CRITICAL PHASE POLYMERIZATIONS Vol. 2

(8)

In addition to the lower molecular weights, the PDIs of the polymers made
in CO2 were broader (3.05.1 vs 1.5). Increasing the monomer concentration by
75% for the same reaction increased the M w of the polymer to 5.9 105 g/mol
and decreased the PDI to 2. This synthesis represents a comprehensible way to
make organometallic polymers from highly strained precursors without dealing
with the limitations imposed by melt-phase (88,89) or polycondensation routes
(90).

Step-Growth Polymerizations in CO2

Melt-Phase Condensation Polymerizations. Many condensation poly-


merizations are performed in the melt phase to produce high molecular weight
material without the need for organic solvents. These polycondensation reactions
are driven by the removal of small molecule condensates such as water, methanol,
and phenol. Current industrial processes require high temperatures and high
vacuum to allow the removal of the condensate molecule from the polymer melt
phase. The drawback to this method is that the high molecular weight polymers
produced have high viscosities that make polymer processing difcult and costly,
as well as limit the molecular weight of the polymer. CO2 is capable of acting as a
plasticizing agent for the polymer melt phase, thus increasing the free volume of
the melt and decreasing the viscosity of the melt for greater processability. Using
CO2 also facilitates the driving force for polycondensation reactions. Plasticiza-
tion of the polymer melt phase provides a greater surface area for condensate
removal, and solubilization of the small molecule condensate in CO2 assists in
carrying it out of the reactor. This enhancement in condensate removal results in
greater mobility of chain ends, leading to higher reaction rates and higher molec-
ular weight products. This strategy has been pioneered by DeSimone to synthe-
size polymers by step-growth methods, including polycarbonates, polyesters, and
polyamides.
Melt polymerization of bisphenols (bisphenol A, bisphenol P, bisphenol AF,
and bisphenol Z) with diphenyl carbonate in CO2 with several catalysts have been
achieved (91,92). Polymers with number-average molecular weights ranging from
2.2 103 to 1.1 104 g/mol (M w = 4.5 103 to 2.7 104 ) were obtained over a
range of reaction temperatures (180250 C) and CO2 pressures [20.724.1 MPa
(207241 bar)]. Reaction conditions were chosen to ensure efcient removal of solu-
bilized condensate (phenol) without extracting the reactants (diphenyl carbonate).
Polycarbonate synthesis from bisphenol A and diphenyl carbonate catalyzed by
tetraphenylphosphonium tetraphenyl borate were also performed in scCO2 (93)
(eq. (9)).
Vol. 2 CRITICAL PHASE POLYMERIZATIONS 123

(9)
Polymers with high molecular weights (M n up to 1.3 104 g/mol) were ob-
tained at 270 C and 29.6 MPa (296 bar) CO2 . In this technique a slight excess of
diphenyl carbonate is required because of monomer extraction. Supercritical CO2
has also been used to induce crystallization of PC prepolymer beads for solid-state
polymerization (94). The amorphous PC beads (M w = 2500 g/mol) were rendered
19% crystalline by treatment with scCO2 and were subsequently polymerized at
temperatures from 180 to 240 C yielding polymers with molecular weights as high
as 1.4 104 g/mol.
Poly(ethylene terephthalate) (PET) is commonly prepared by melt-phase
polycondensation reactions. Bis(hydroxyethyl) terephthalate (BHET) was con-
verted to PET under a variable ow (210 mL/min) of CO2 at 20.7 MPa (207 bar)
(95). Molecular weights were obtained over the range 3 103 to 6.3 103 g/mol and
increased signicantly with ow rate and/or reaction time. These values are low
compared to normal molecular weights obtained in commercial processes (approx.
2 104 g/mol) and may be attributed to the decreased solubility of the ethylene
glycol condensate in CO2 compared with phenol in the polycarbonate system.
Polyamides have also been synthesized in the melt phase in the presence of
CO2 by the nylon salt route (96). A 1:1 salt of hexamethylenediamine and adipic
acid was heated at 220 C for 2 h and then at 280 C for 3 h in 20.7 MPa (207 bar)
CO2 producing nylon-6,6 with molecular weights up to 2.45 104 g/mol.
Oxidative Coupling. The synthesis of polypyrrole, an organic conducting
polymer known for its good thermal stability, has been extended to SFCs. Polypyr-
role was synthesized in supercritical uoroform and in scCO2 from pyrrole that is
generated in situ by decarboxylation of 2-carboxypyrrole and a chemical oxidant
[either FeCl3 or Fe(CF3 SO3 )3 ] (97). Fe(CF3 SO3 )3 was more soluble in CO2 and
gave higher yields compared to FeCl3 (5060% vs. 87% yield). These yields are still
lower than those obtained from control reactions run in toluene (92%). Unlike the
globular structure normally obtained when polypyrrole is prepared in traditional
solvents, these polymers possess a different morphology with bers ranging from
100 to 200 nm in diameter and several micrometers long.
Poly(2,6-dimethylphenylene oxide) (PPO) is synthesized via oxidative cou-
pling in a CO2 continuous phase (43). The reactions proceed at 34.5 MPa (345 bar)
and at either room temperature or 40 C for 20 h using a CuBr/amine/O2 catalyst
system (eq. (10)). PPO with yields as high as 83% and Mn up to 1.7 104 g/mol were
precipitated. Small molecule amines (pyridine and dimethylethylamine) and CO2 -
soluble polymeric amines (block copolymers of FOA and either 4-vinylpyridine or
DMAEA) were used in the reaction.
124 CRITICAL PHASE POLYMERIZATIONS Vol. 2

(10)
The pyridine proved to be the most effective amine for the system. Although
the polymeric amines produced milky white latexes and served to stabilize the
PPO as a colloid, no increase in yields or molecular weight associated with disper-
sion polymerizations were observed.
Porous Polymer Synthesis. Supercritical CO2 is an attractive solvent
for the preparation of porous polymers since conventional processes often require
large volumes of porogenic organic solvents that tend to become trapped in
the polymer matrix. Microcellular polyurethane foams have been synthesized in
scCO2 by treating a range of diisocyanates with various propylene oxide and ethy-
lene oxide polyols (98). Phase separation was found to lead to the generation of
a polymer-rich phase and a CO2 -rich phase that became porous upon removal
of CO2 . Another example of the synthesis of porous polymers is the direct for-
mation of aerogels by solgel polymerization of alkoxysilanes in scCO2 (99). A
water-free solgel polymerization technique is utilized in order to account for
the low miscibility of water and CO2 (100). In this technique, alkoxysilanes such
as tetramethoxysilane (TMOS) and 1,4-bis-(triethoxysilyl) benzene (BESP) can
react with anhydrous formic acid at 3540 C in the presence of 41 MPa (410 bar)
scCO2 for up to 12 h. Opaque white silica aerogels were obtained in nearly quanti-
tative yields (after aging for 1218 h). The silica gels contained pores larger than
those found in gels made in traditional solvents and had interconnected particles
100200 nm in diameter. Surface areas of silica gels (260308 m2 /g) were similar
to gels formed in ethanol or formic acid. BESP aerogels had surface areas lower
than those prepared by aqueous solgel polymerizations in organic solvents (471
586 m2 /g vs 10001600 m2 /g) (101). A technique has been developed that exploits
the selation of CO2 to form well-dened macroporous materials based on acrylate
monomers (IOZ) and by using suspension polymerization methods (103).

Conclusion

The many examples cited in this article suggest that CO2 is a viable alternative
solvent for numerous polymerizations. The use of CO2 provides not only environ-
mental advantages over traditional solvents but also performance advantages.
And its tunable properties allow it to be used as a continuous phase in a vari-
ety of both step-growth and chain-growth polymerization techniques with con-
trol heretofore unattainable with conventional solvents. The rapid growth of CO2
technology in the past ten years has provided fundamental knowledge that will
undoubtedly enable the preparation of more advanced materials in the future.

BIBLIOGRAPHY

1. J. A. Darr and M. Poliakoff, Chem. Rev. 99, 565 (1999).


2. K. M. Scholsky, J. Supercrit. Fluids 6, 103 (1993).
3. A. I. Cooper, J. Mater. Chem. 10, 207 (2000); J. L. Kendall and co-workers, Chem.
Rev. 99, 543 (1999).
Vol. 2 CRITICAL PHASE POLYMERIZATIONS 125

4. E. L. Quinn and C. L. Jones, Carbon Dioxide, Reinhold Publishing Corp., New York,
1936.
5. J. A. Hyatt, J. Org. Chem. 49, 5097 (1984).
6. M. A. McHugh and V. J. Krukonis, Supercritical Fluid Extraction: Principles and
Practice, 2nd ed., Butterworth-Heineman, Stoneham, 1993.
7. P. E. Jessop and W. Leitner, eds., Chemical Synthesis Using Supercritical Fluids,
Wiley-VCH, Weinheim, 1999.
8. T. Sarbu, T. Styranec, and E. J. Beckman, Nature 405, 165 (2000).
9. R. C. Reid, J. M. Praunsnitz, and B. E. Poling, The Properties of Gases and Liquids,
4th ed., McGraw-Hill Book Co., Inc., New York, 1987.
10. F. G. Keyes and J. G. Kirkwood, Phys. Rev. 36, 754 (1930).
11. S. G. Kazarian and co-workers, J. Am. Chem. Soc. 118, 1729 (1996).
12. A. Dardin and co-workers, Macromolecules 30, 3593 (1997).
13. A. Dardin, J. M. DeSimone, and E. T. Samulski, J. Phys. Chem. B 102, 1775 (1998).
14. J. M. DeSimone, Z. Guan, and C. S. Elsbernd, Science 257, 945 (1992).
15. Z. Guan and co-workers, Macromolecules 26, 2663 (1993).
16. Int. Pat. Appl. PCT/US 93/01626 (1993), J. M. DeSimone.
17. Z. Guan and co-workers, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 34, 329
(1993).
18. K. K. Kapellen, C. D. Mistele, and J. M. DeSimone, Polym. Prepr. (Am. Chem. Soc. Div.
Polym. Mater. Sci. Eng.) 74, 256 (1996).
19. M. Shaw, J. Ryan, and C. Erkey, Abstr. Pap. Am. Chem. Soc. 214, 243 (1997).
20. A. M. van Herk and co-workers, Macromolecules 30, 4780 (1997).
21. M. A. Quadir and co-workers, Macromolecules 31, 6481 (1998).
22. S. Beuermann and co-workers, Macromol. Chem. Phys. 199, 1209 (1998).
23. J. R. Combes, Z. Guan, and J. M. DeSimone, Macromolecules 27, 865 (1994).
24. M. Hagiwara and co-workers, J. Polym. Sci., Part A-1 6, 603 (1968).
25. M. Hagiwara and co-workers, J. Polym. Sci., Part A-1 6, 609 (1968).
26. U.S. Pat. 3471463 (1969), T. Kagiya and co-workers.
27. Fr. Pat. 1524533 (1968) (Sumitomo Chemical Co.).
28. U.S. Pat. 3522228 (1970), K. Fukui and co-workers.
29. T. J. Romack, E. E. Maury, and J. M. DeSimone, Macromolecules 28, 912 (1995).
30. H. Q. Hu and co-workers, Acta Polym. Sinica 740 (1998).
31. U.S. Pat. 5345013 (1994), D. J. Van Bramer, M. B. Shiett, and A. Yokozeki.
32. T. J. Romack, B. E. Kipp, and J. M. DeSimone, Macromolecules 28, 8432 (1995).
33. T. J. Romack and J. M. DeSimone, Macromolecules 28, 8429 (1995).
34. J. P. DeYoung, T. J. Romack, and J. M. DeSimone, Polym. Prepr. (Am. Chem. Soc., Div.
Polym. Chem.) 38, 424 (1997).
35. P. A. Carpentier and co-workers, Macromolecules 32, 5973 (1999).
36. A. I. Cooper, W. P. Hems, and A. B. Holmes, Macromol. Rapid Commun. 19, 353 (1998).
37. A. I. Cooper, W. P. Hems, and A. B. Holmes, Macromolecules 32, 2156 (1999).
38. K. E. J. Barret, Dispersion Polymerization in Organic Media, John Wiley & Sons Inc.,
New York, 1975
39. E. D. Sudol, in J. M. Asua, ed., Polymeric Dispersions: Principles and Applications,
Vol. 335, Kluwer Academic Press: Boston, Mass. 1997, pp. 141153.
40. J. M. DeSimone and co-workers, Science 265, 356 (1994).
41. F. A. Adamsky and E. J. Beckman, Macromolecules 27, 312 (1994).
42. K. K. Kapellen, C. D. Mistelle, and J. M. DeSimone, Macromolecules 29, 495 (1996).
43. Y. L. Hsiao and co-workers, Macromolecules 28, 8159 (1995).
44. Y. L. Hsiao and J. M. DeSimone, J. Polym. Sci., Part A: Polym. Chem. 35, 2009 (1997).
45. K. A. Shaffer and co-workers, Macromolecules 29, 2704 (1996).
46. M. L. ONeill and co-workers, Macromolecules 31, 2848 (1998).
47. M. L. ONeill and co-workers, Macromolecules 31, 2838 (1998).
126 CRITICAL PHASE POLYMERIZATIONS Vol. 2

48. C. Lepilleur and E. J. Beckman, Macromolecules 30, 745 (1997).


49. D. A. Canelas, D. E. Betts, and J. M. DeSimone, Macromolecules 29, 2818 (1996).
50. D. A. Canelas and J. M. DeSimone, Macromolecules 30, 5673 (1997).
51. M. R. Clarke, J. L. Kendall, and J. M. DeSimone, Macromolecules 30, 6011 (1997).
52. P. Christian and S. M. Howdle, Macromolecules 33, 237 (2000).
53. D. A. Canelas and co-workers, Macromolecules 31, 6794 (1998).
54. T. Carson, J. Lizotte, and J. M. DeSimone, Macromolecules, 33, 19171920 2000.
55. H. Shiho and J. M. DeSimone, Macromolecules, 33, 15651569 2000.
56. D. C. Blackley, Emulsion Polymerization, John Wiley & Sons, Inc., New York, 1975.
57. R. G. Gilbert, Emulsion Polymerization: A Mechanistic Approach, Academic Press,
London, 1995.
58. F. A. Adamsky and E. J. Beckman, Macromolecules 27, 523 (1994).
59. E. E. Maury and co-workers, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 6,
205 (1993).
60. K. P. Johnston and co-workers, Science 271, 624 (1996).
61. K. Harrison and co-workers, Langmuir 10, 3536 (1994).
62. J. J. Watkins and T. J. McCarthy, Macromolecules 27, 4845 (1994).
63. J. J. Watkins and T. J. McCarthy, Macromolecules 28, 4067 (1995).
64. E. Kung, A. J. Lesser, and T. J. McCarthy, Macromolecules 31, 4160 (1997).
65. K. A. Arora, A. J. Lesser, and T. J. McCarthy, Macromolecules 32, 2562 (1999).
66. P. Rajagopalan and T. J. McCarthy, Macromolecules 31, 4791 (1998).
67. R. Biddulph and P. Plesch, J. Chem. Soc. 3913 (1960).
68. T. Perneker and J. P. Kennedy, Polym. Bull. 32, 537 (1994).
69. U.S. Pat. 5376744 (1994), J. P. Kennedy and T. Pernecker.
70. H. Yokota and co-workers, J. Polym. Sci., Part A-1 42, 1412 (1969).
71. T. Pernecker, G. Deak, and J. P. Kennedy, Macromol. Rep. A 32, 969 (1995).
72. M. R. Clarke and J. M. DeSimone, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.)
35, 482 (1994).
73. M. R. Clarke and J. M. DeSimone, Macromolecules 28, 3002 (1995).
74. T. Higashimura, M. Sawamoto, and M. Miyamoto, Macromolecules 17, 265 (1984).
75. M. R. Clarke, J. L. Kendall, and J. M. DeSimone, Macromolecules 30, 6011 (1997).
76. S. T. Nguyen, R. H. Grubbs, and J. W. Ziller, J. Am. Chem. Soc. 115, 9858 (1993).
77. C. D. Mistle, H. H. Thorp, and J. M. DeSimone, J. Macromol. Sci., A: Pure Appl. Chem.
33, 953 (1996).
78. J. G. Hamilton and co-workers, Macromolecules 31, 4387 (1998).
79. P. Schwab and co-workers, Angew. Chem. Int. Ed. Engl. 34, 2039 (1995).
80. P. Schwab, R. H. Grubbs, and J. W. Ziller, J. Am. Chem. Soc. 118, 100 (1996).
81. R. R. Schrock and co-workers, J. Am Chem. Soc. 112, 3875 (1990).
82. H. Hori, C. Six, and W. Lietner, Macromolecules 32, 3178 (1999).
83. D. J. Darensbourg, N. W. Stafford, and T. Katsurao, J. Mol. Cat. A: Chem. 104, L1L4
(1995).
84. D. J. Darensbourg and M. W. Holtcamp, Macromolecules 28, 7577 (1995).
85. D. J. Darnsbourg and M. W. Holtcamp, Coord. Chem. Rev. 153, 155 (1996).
86. S. A. Mang and co-workers, Macromolecules, in press.
87. P. Christrian and co-workers, Polym. Prepr. 8, 434 (1997).
88. C. U. Pittman and S. Yukihiko, Chem. Lett. 383 (1975).
89. D. A. Foucher, B. Z. Tang, and I. Manners, J. Am. Chem. Soc. 114, 6246 (1992).
90. E. W. Neuse and L. Bednarik, Macromolecules 12, 187 (1979).
91. P. G. Odell and G. K. Hamer, Polym. Prepr. 38, 470 (1997).
92. U.S. Pat. 5698665 (1997), P. G. Odell.
93. S. M. Gross and co-workers, Macromolecules 31, 25, 9090 (1998).
94. S. M. Gross and co-workers, Macromolecules 32, 3167 (1999).
95. A. L. Burke, G. Maier, and J. M. DeSimone, Polym. Mater. Sci. Eng. 74, 248 (1996).

Das könnte Ihnen auch gefallen