[13]desimone

download [13]desimone

of 7

Transcript of [13]desimone

  • 7/28/2019 [13]desimone

    1/7

    Green synthesis of polymers using supercritical carbon dioxide

    Colin D. Wood a, Andrew I. Cooper b,*, Joseph M. DeSimone a,c,*

    a Department of Chemistry, University of North Carolina at Chapel Hill, NC 27599, USAb Donnan and Robert Robinson Laboratories, Department of Chemistry, University of Liverpool, Liverpool, L69 3BX, UK

    c Department of Chemical Engineering, North Carolina State University, Raleigh, NC 27695, USA

    Abstract

    Supercritical carbon dioxide is the most extensively studied supercritical fluid (SCF) medium for polymerization reactions and

    organic transformations. This can be attributed to a list of advantages ranging from solvent properties to practical environmental

    as well as economic considerations. Aside from these gains, CO2 finds particularly advantageous application in the synthesis and

    processing of fluoropolymers and porous materials and as such, this will be the subject of this review. The overall aim is to highlight

    areas where the unique properties of SCF solvents can be exploited to generate polymers that would be difficult or inconvenient to

    obtain by other routes.

    2005 Elsevier Ltd. All rights reserved.

    Keywords: Supercritical; Fluoropolymer; Porous polymers; High-throughput

    1. Introduction

    Supercritical carbon dioxide has emerged as the most

    extensively studied environmentally benign medium for

    organic transformations and polymerization reactions.

    This stems from a list of advantages ranging from sol-

    vent properties to practical environmental and economic

    considerations. Increasing environmental regulation has

    prompted a tremendous effort to develop and apply

    environmentally benign alternative media or solvent free

    processes to a variety of commercially significant, sol-

    vent-intensive processes [1]. Carbon dioxide in the liquid

    and supercritical state is a potential alternative for manyof the processes experiencing ever increasing regulation,

    owing to its non-toxic, non-flammable nature. Addition-

    ally, its ubiquity makes it inexpensive and readily avail-

    able. Several excellent reviews are available that discuss

    the utilization of carbon dioxide for many solvent- and

    waste-intensive industries [25]. Aside from the gains

    provided by CO2 as a reaction medium in general, itfinds particularly advantageous application in the syn-

    thesis and processing of fluoropolymers and porous

    materials [611] as such, this will be the subject of this

    review. In particular the aim is to highlight areas where

    the unique properties of supercritical fluid (SCF) sol-

    vents can be exploited to generate polymers that would

    be difficult or inconvenient to obtain by other routes.

    The general properties of SCFs in relation to chemical

    synthesis [8] and extraction [2] have been reviewed previ-

    ously and will not be a focus herein.

    2. Fluoroolefin polymerization in CO2

    2.1. Introduction

    Fluoropolymers are typically synthesized in aqueous

    polymerization systems (both emulsion and suspension),

    non-aqueous systems (Freon-113), or in hybrid Freon-

    113/aqueous hybrid systems [12]. Such processes require

    the use of large quantities of water, CFCs (for non-aqueous

    1359-0286/$ - see front matter 2005 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.cossms.2005.02.001

    * Corresponding authors.

    E-mail address: [email protected] (A.I. Cooper).

    Current Opinion in Solid State and Materials Science 8 (2004) 325331

    mailto:[email protected]:[email protected]
  • 7/28/2019 [13]desimone

    2/7

    polymerizations), and fluorinated surfactants for emul-

    sion polymerization. Many of the fluorinated surfac-

    tants typically employed in aqueous emulsion and

    suspension polymerizations are currently under scrutiny

    due to bioaccumulation and environmental persistence

    [13,14]. These issues have collectively resulted in an

    increasingly urgent impetus for transition from the con-ventional fluoropolymer synthesis platforms to alterna-

    tives that meet the requirements of emerging public

    and regulatory demands. Most commercially available

    fluoropolymers are prepared from a relatively small

    group of olefins including tetrafluoroethylene (TFE),

    chlorotrifluoroethylene (CTFE), vinylidene fluoride

    (VF2), hexafluoropropylene (HFP), ethylene, and per-

    fluoroalkyl vinyl ethers (PAVEs). Many of these mono-

    mers are flammable and some are explosive. For

    example, TFE is flammable when mixed with air and

    has a high propensity for explosion during expansion

    to a gas from its liquid phase under pressure. Further,

    TFE is highly explosive as a gas at elevated tempera-

    tures. In the presence of oxygen it will undergo autopo-

    lymerization, a process sufficiently exothermic to ignite

    an explosion. Aqueous suspension and dispersion poly-

    merizations of copolymers of TFE with a variety of

    comonomers including hexafluoropropylene (HFP)

    and various perfluoroalkyl vinyl ethers (PAVEs) typi-

    cally exhibit high levels of carboxylic acid end groups

    [7,15]. The presence of acid end groups often proves del-

    eterious to the intended properties and function of the

    polymer not to mention the introduction of complica-

    tions into certain post polymerization, melt-processing

    steps. This problem is particularly prevalent in copoly-mers including PAVE comonomers (perfluoroalkyl vinyl

    ethers) which are an essential component in the control

    of crystallinity in TFE-based fluoroplastics. In order to

    prevent decomposition, discoloration and emissions of

    hydrogen fluoride, polymers containing high levels of

    acid end groups may require high-temperature hydroly-

    sis and fluorination finishing steps. A propagating poly-

    meric radical derived from a terminal PAVE monomer

    unit has one of two possible reaction pathways (Scheme

    1). First, and most obvious, it can cross propagate to

    monomer continuing the polymerization reaction or itcan undergo b-scission, resulting in an acid fluoride ter-

    minated polymer and a perfluoroalkyl radical capable of

    initiating further polymerization which is essentially a

    chain-transfer-to-monomer step [7,15]. The use of CFCs

    circumvents these problems; however, CFCs have fallen

    under exceedingly strict regulation due to environmental

    concerns and as a result, are no longer viable to be used

    as large scale reaction media.

    2.2. TFE-based materials

    Tetrafluoroethylene (TFE)-based copolymers have

    become premium high performance materials for a

    broad range of applications that demand chemical and

    thermal resistance, and melt processing capability [16].

    As mentioned earlier, caution must be taken when

    TFE is used as a monomer due to potential explosions.

    It has been demonstrated that CO2/TFE mixtures are far

    less susceptible to ignition as TFE forms a pseudo

    azeotrope with CO2 [17]. This makes handling and deliv-

    ering monomer much safer. A range of TFE-based

    fluoropolymers have been successfully synthesized in

    CO2. These include FEP, PFA, ETFE, TFE/vinyl ace-

    tate polymers, Nafion-type materials, and Teflon

    AF-type materials (Fig. 1) the details have recently beenreviewed elsewhere [11]. Further, it is worth noting that

    fluoropolymers synthesized in sc-CO2 exhibit signifi-

    cantly diminished levels of acid end groups leading to

    very high molecular weight materials [15]. Indeed, in

    such instances, it is even necessary to add chain transfer

    agents to control molecular weight and maintain melt-

    processability of the product [18]. IR analysis has con-

    firmed that fluoropolymers synthesized under these

    conditions exhibit acid end group levels an order of

    magnitude below (03 end groups per 106 carbon atoms)

    those which are synthesized in conventional organic and

    aqueous reaction systems [14,18]. The most likely expla-

    nation for this is that the tremendous plasticizing

    capability that CO2 provides for efficient transport of

    monomer into the polymer phase, maintaining high

    effective concentrations of monomer in the vicinity of

    the active chain ends, thus favoring the bimolecular

    propagation step over the unimolecular b-scission pro-

    cess (Scheme 1). The other explanation may be the lower

    temperatures that one can use with CO2 which may

    favor cross propagation over b-scission. Additionally,

    it has been demonstrated that dense CO2 is exceptionally

    inert to free radical chemistry in general [19]. This dem-

    onstrates the broad applicability of liquid and supercrit-

    CF2 C

    F

    O

    CF2

    CF2

    CF3

    CF2 C

    F

    O

    CF2

    CF2

    CF3

    CF2 C

    F

    F

    CF2 C

    F

    O

    CF2

    CF2

    CF3

    CF2 C

    O

    F

    F3C CF2 CF2

    TFE

    CrossPropagation

    +

    -scission

    Scheme 1. Reaction sequence for b-scission in fluoroolefin polymer-

    ization [7,15].

    326 C.D. Wood et al. / Current Opinion in Solid State and Materials Science 8 (2004) 325331

  • 7/28/2019 [13]desimone

    3/7

    ical CO2 in the synthesis and processing of fluorinated

    polymers which has led to the commercialization of a

    CO2-based approach in the manufacturing of certain

    grades of polymers based on tetrafluoroethylene (Tef-

    lonTM). DuPont has recently invested an initial $40 mil-

    lion in the construction of a TeflonTM FEP production

    facility in North Carolina. The process employs CO2as the continuous phase instead of 1,1,2-trichloro-

    1,2,2-trifluoroethane or water and surfactant [1,2023].

    The use of carbon dioxide results in a surfactant free

    system with no issues of chain-transfer-to-solvent. This

    process not only avoids many of the environmental

    and safety concerns but also provides a product with

    superior properties. This is the first commercial example

    of fluoropolymer resins made using supercritical carbon

    dioxide as the solvent.

    Teflon AF is an amorphous copolymer of tetraflu-

    oroethylene (TFE) and 2,2-bis(trifluoromethyl)-4,5-

    difluoro-1,3-dioxole (PDD). It combines the properties

    of amorphous plastics, such as optical transparency

    and solubility in organic solvents, with those of perfluo-

    rinated polymers, including high thermal stability, excel-

    lent chemical stability and low surface energy. Moreover

    Teflon AF exhibits the lowest dielectric constant (1.90

    for Teflon AF 2400) and the lowest refractive index

    (1.29 for Teflon AF 2400) known for any solid organic

    polymer [12]. As such, Teflon AF is well-suited for use

    as an optical material. Teflon AF-based materials have

    also been synthesized in supercritical carbon dioxide

    (see Fig. 1) [24]. A range of copolymers with various

    compositions and different molecular weights were pre-

    pared in yields as high as 74% and their properties were

    compared to commercially available Teflon AF. The

    glass transition temperatures for the materials ranged

    from 67 to 334 C. The synthesis of these copolymers

    in CO2 has several key advantages compared to conven-

    tional polymerization techniques. The low reaction tem-

    perature and the use of a perfluorinated initiator resultsin copolymers with properties which are comparable to

    those of the analogous commercial product. However,

    an additional fluorination step is not necessary and the

    product is isolated directly from the reactor without

    contamination from solvents or surfactants. Moreover,

    the synthesis uses TFE/CO2 mixtures instead of pure

    TFE which has inherent safety advantages [17].

    2.3. VF2-based materials

    VF2 has been successfully synthesized in CO2 in a

    continuous precipitation process using a continuous

    stirred tank reactor [25]. In such a process monomer,

    solvent (CO2), and initiator are continuously added to

    the reactor. The advantage of a continuous process over

    a batch process is that the monomer and the CO2 can be

    removed and recycled, leading to a high rate of polymer-

    ization. The authors later reported that, although this

    was a precipitation polymerization, a model based on

    the kinetics for a homogeneous polymerization fit the

    experimental data, confirming the reaction order of 0.5

    and 1 for the initiator and monomer respectively [26].

    Further, the effect of VF2 flow rate, residence time

    and temperature were investigated in this system [27].

    Fluorinated Ethylene Propylene Resin

    (FEP)

    Perfluoroalkoxy Resin

    (PFA)

    Ethylene Tetrafluoroethylene Resin

    (ETFE)

    Nafion

    CF2 CF2 CH2 CH2

    CF2 CF2 CF2 CF

    ORf

    CF2 CF2 CF2 CF

    CF3

    CF2 CF2 CF2 CF

    OCF2CF(CF3)OCF2CF2SO2F

    CFCF

    O O

    CF3F3C

    CF2CF2

    n

    nn

    n n

    Teflon

    AF

    CF2 CF2 CH2 CH

    n

    O

    O

    CH3

    Tetrafluoroethylene-co-vinyl acetate

    Fig. 1. Tetrafluoroethylene-based fluoropolymers synthesized in carbon dioxide.

    C.D. Wood et al. / Current Opinion in Solid State and Materials Science 8 (2004) 325331 327

  • 7/28/2019 [13]desimone

    4/7

    It was found that monomer feed concentration had a

    pronounced effect upon the modality of the molecular

    weight distribution. At a high monomer concentration,

    the resulting polymer exhibited a bimodal molecular

    weight distribution. The authors suggested that at high

    monomer concentrations the kinetics may no longer ad-

    here to the homogeneous kinetic model proposed earlier[26]. Several other parameters were investigated such as

    chain branching and agitation but they were unable to

    explain the bimodality of the polymer.

    2.4. Fluoroalkyl acrylate polymerizations in CO2

    One of the more difficult issues associated with the

    use of CO2 as a polymerization medium has been the

    limited solubility of most high molecular weight poly-

    mers in CO2. As such, much of the work reported for

    polymerizations in CO2 are typically heterogeneous pre-

    cipitation, dispersion or emulsion processes as discussed

    above. However, it has been demonstrated that oligo-

    meric perfluoropolyethers and oligomeric poly(chlo-

    rotrifluoroethylene) oils are soluble in liquid CO2[2]. Additionally, it has been reported that highly fluori-

    nated polyacrylates of high molecular weight (>250,000)

    exhibit exceptional solubility in supercritical CO2. Not

    surprisingly, since this discovery, it has been demon-

    strated that high molecular weight, amorphous fluoro-

    polymers can be synthesized homogeneously in CO2utilizing free radical initiators [21,2833]. Most amor-

    phous fluoropolymers exhibit solvent resistance to

    common organic solvents and therefore conventional

    approaches to synthesis and processing have dependedon chlorofluorocarbon (CFC) platforms. Owing to the

    previously mentioned environmental concerns with the

    use of CFC refrigerants and solvents, CFCs are no

    longer acceptable for large-scale commercial use. As

    an effective reaction medium for homogeneous polymer-

    izations under mild conditions, CO2, again, is an

    increasingly attractive, inexpensive and harmless plat-

    form option for the synthesis of highly fluorinated

    monomers (as domestic and international regulation of

    CFCs becomes ever more restrictive in the coming

    years). Various fluorinated acrylate monomers, such as

    1,1-dihydroperfluorooctyl acrylate (FOA) have been

    successfully polymerized in CO2. It is also noted that

    reaction conditions for these polymerizations are mild

    in terms of temperature and pressure and not entirely

    unlike those encountered in conventional solvent based

    processes [21]. Further, studies have shown CO2 to be

    inert to radical chemistry, eliminating chain transfer to

    solvent as a side reaction [8]. Additionally the low vis-

    cosity of sc-CO2 removes issues of Tromsdorf effect

    and autoacceleration [8]. Therefore, as a reaction med-

    ium, CO2 is ideally suited to free radical polymeriza-

    tions. In all cases, the polymerizations remained

    homogeneous throughout the reaction, illustrating the

    high solubility of the polymers in supercritical CO2.

    The solution properties of poly(FOA) in CO2 have also

    been investigated using small-angle neutron scattering

    (SANS) and it was shown that CO2 is an excellent sol-

    vent for poly(FOA) [34].

    2.5. Photooxidation of fluoroolefins in liquid CO2

    Another class of fluoropolymer known to be readily

    soluble in CO2 are perfluoropolyethers (PFPEs) [2]. As

    a unique class of fluoropolymers, PFPE polymers and

    copolymers have been established as high performance

    materials, exhibiting low surface energies and low

    moduli, as well as excellent thermal and chemical stabil-

    ities. PFPEs are primarily found in high performance

    lubricant applications such as for magnetic data storage

    media and as heat exchanger fluids. Recently, DeSimone

    et al. developed a novel solvent resistant photocurable

    PFPE that can be used for the fabrication of microflu-

    idic devices [35] and as materials for high resolution soft

    lithography to replicate sub-100 nm sized features with

    no indications of limits to going to even smaller in size

    [36]. Moreover, the material has several specific advan-

    tages over traditional glass and PDMS-based materials

    and broadens the utility of both processes. In addition,

    the process exemplifies a solvent-free approach.

    One of the main industrial processes for the produc-

    tion of PFPEs is photooxidation of fluoroolefins [37].

    Currently, only TFE and HFP are used commercially

    in this process. Typically, HFP is photooxidized in bulk

    owing to its very low reactivity while TFE requires an

    inert diluent in order to prevent homopolymerizationof the olefin. DeSimone and coworkers reported recently

    the photooxidation of various concentrations of HFP

    in CO2 in concert with parallel reactions carried out in

    bulk HFP and in perfluorocyclobutane for purposes of

    comparison [38]. Based on the data collected in this

    study, it was demonstrated that there is a strong depen-

    dence of molecular weight and composition on HFP

    concentration (i.e., lower HFP concentrations gave

    lower peroxide content). Additionally, viscosity effects

    in liquid CO2 appeared to significantly reduce the

    amount of peroxidic linkages in the final product.

    2.6. Porous materials and supercritical fluids

    Porous materials are used in a wide range of appli-

    cations, including catalysis, chemical separations, and

    tissue engineering [9]. However, the synthesis of these

    materials is often solvent intensive. Supercritical carbon

    dioxide as an alternative solvent for the production of

    functional porous materials can circumvent this issue

    as well as affording a number of specific physical, chem-

    ical, and toxicological advantages listed below. For

    example, in order to dry porous materials often requires

    energy intensive steps whereas; the transient dry nat-

    328 C.D. Wood et al. / Current Opinion in Solid State and Materials Science 8 (2004) 325331

  • 7/28/2019 [13]desimone

    5/7

    ure of CO2 circumvents these issues. Pore collapse can

    occur in certain materials when removing conventional

    liquid solvents; this can be avoided using SCFs because

    they do not possess a liquidvapour interface. Porous

    structures are important in biomedical applications

    (e.g., tissue engineering) where the low toxicity of CO2

    offers specific advantages in terms of minimizing theuse of toxic organic solvents. In addition, the wetting

    properties and low viscosity of CO2 offers specific bene-

    fits in terms of surface modification.

    2.7. Free-radical polymerization using sc-CO2 as a

    pressure-adjustable porogen

    Recently, Cooper et al. demonstrated the formation

    of permanently porous crosslinked poly(acrylate) and

    poly(methacrylates) monoliths using sc-CO2 as the por-

    ogenic solvent [39,40]. Materials of this type [41,42] are

    useful in applications such as high-performance liquid

    chromatography, high-performance membrane chroma-

    tography, capillary electrochromatography, microflui-

    dics, [43] and high-throughput bioreactors [44]. In this

    process, no organic solvents are used in either the syn-

    thesis or purification. It is possible to synthesize the

    monoliths in a variety of containment vessels (Fig. 2),

    including chromatography columns and narrow-bore

    capillary tubing. Moreover, the variable density associ-

    ated with SCF solvents was exploited in order to

    fine-tune the polymer morphology. Fig. 3 shows the

    variation in BrunnauerEmmetTeller (BET) surface

    area for a series of crosslinked monoliths synthesized

    using sc-CO2 as the porogen over a range of reactionpressures [45]. As can be seen the BET surface area

    and the average pore size in these materials could be

    continuously varied over a broad range simply by vary-

    ing the SCF solvent density. It is interesting to note the

    minimum in area (and a maximum in the average pore

    diameter, not shown) was observed at a reaction pres-

    sure of around 180 bar. This can be rationalized by

    considering the variation in solvent quality as a function

    of CO2 density and the resulting influence on the mech-

    anism of nucleation, phase separation, aggregation,

    monomer partitioning, and pore formation [46]. This

    approach was also extended to synthesize well-defined

    porous, cross-linked beads by suspension polymeriza-

    tion, again without using any organic solvents [47].

    The surface area of the of the beads could be tuned over

    a broad range (5500 m2/g) simply by varying the CO2density.

    2.8. Templating of supercritical fluid emulsions

    Emulsion templating is useful for the synthesis ofhighly porous inorganic, [4851] and organic materials

    [5254]. In principle, it is possible to access a wide range

    of porous hydrophilic materials by reaction-induced

    phase separation of concentrated oil-in-water (O/W)

    emulsions. A significant drawback to this process is that

    large quantities of a water-immiscible oil or organic

    solvent are required as the internal phase (usually

    >80 vol.%). In addition, it is often difficult to remove

    this oil phase after the reaction. Inspired by studies on

    SCF emulsion stability and formation [55], we have

    developed methods for templating high internal phase

    CO2-in-water (C/W) emulsions (HIPE) to generate

    highly porous materials in the absence of any organic

    solventsonly water and CO2 are used [56]. If the emul-

    sions are stable one can generate low-density materials

    (0.1 g/cm3) with relatively large pore volumes (up to

    6 cm3/g) from water-soluble vinyl monomers such as

    acrylamide and hydroxyethyl acrylate. Fig. 4 shows a

    crosslinked polyacrylamide material synthesized from

    a high internal phase C/W emulsion, as characterized

    by SEM and confocal microscopy (scale = 230 lm

    230 lm). Comparison of the two images illustrates quite

    clearly how the porous structure shown in the SEM

    image is templated from the C/W emulsion (as

    Fig. 2. Polymer monolith, synthesized using sc-CO2 as the porogen as

    isolated from the reaction vessel. The monolith adopts the shape of the

    reactor. Reproduced with permission; copyright 2003, American

    Chemical Society [45].

    50

    100

    150

    200

    250

    300

    350

    100 150 200 2 300 350 4 45

    Pressure (bar)

    250 300 400 450

    Surfac

    eArea(m2/g)

    Fig. 3. Surface area (as measured by N2 adsorptiondesorption using

    the BET method) as a function of the CO2 pressure employed during

    the polymerization. Reproduced with permission; copyright 2003,

    American Chemical Society [45].

    C.D. Wood et al. / Current Opinion in Solid State and Materials Science 8 (2004) 325331 329

  • 7/28/2019 [13]desimone

    6/7

    represented by the confocal microscopy image of the

    pores). In general, the confocal image gives a better mea-

    sure of the CO2 emulsion droplet size and size distribu-

    tion immediately before gelation of the aqueous phase.

    Initially we used low molecular weight (Mw 550 g/

    mol) perfluoropolyether ammonium carboxylate surfac-

    tants to stabilize the C/W emulsions [56] but there are

    some practical disadvantages of using these surfactants

    in this particular process such as cost and the surfactant

    is non-degradable. It was subsequently shown that it is

    possible to use inexpensive hydrocarbon surfactants to

    stabilize C/W emulsions and that these emulsions can

    also be templated to yield low-density porous materials

    [57]. In this study it was shown that all of the problems

    associated with the initial approach could be overcome

    and it was possible to synthesize C/W emulsion-tem-

    plated polymers at relatively modest pressures (60

    70 bar) and low temperatures (20 C) using inexpensive

    and readily available hydrocarbon surfactants. More-

    over, we demonstrated that this technique can in princi-ple be extended to the synthesis of emulsion-templated

    HEA and HEMA hydrogels that may be useful, for

    example, in biomedical applications [5860].

    2.9. High throughput solubility measurements in CO2

    As discussed CO2 is an excellent environmentally

    benign alternative solvent for the synthesis of fluoro-

    polymers and porous materials. Moreover, a range of

    environmental as well as practical advantages can be

    realized from using CO2, which has resulted in the com-

    mercial synthesis of certain grades of Teflon by DuPont.

    However, the discovery of inexpensive CO2-soluble sur-

    factants, ligands and phase transfer catalysts would

    greatly increase the utility of CO2. As mentioned the

    only polymers with appreciable solubility in CO2 are

    amorphous fluoropolymers and polysiloxanes [61]. Inex-

    pensive poly(ether carbonate) copolymers have been re-

    ported to be soluble in CO2 under moderate conditions

    [62,63] and could function as building blocks for inex-

    pensive surfactants, but numerous practical difficulties

    remain. Recently, Cooper et al. reported a new method

    which allows for the rapid parallel solubility measure-

    ments for libraries of materials in supercritical fluids

    [64]. The technique was used to evaluate the solubilityof a mixed library of 100 synthetic polymers including

    polyesters, polycarbonates, and vinyl polymers. This

    method is at least 50 times faster than other techniques

    in terms of the rate of useful information that is ob-

    tained and has broad utility in the discovery of novel

    SCF-soluble ligands, catalysts, biomolecules, dyes, or

    pharmaceuticals for a wide range of materials

    applications.

    3. Conclusions

    In general CO2 is an attractive solvent alternative for

    the synthesis of polymers because it is environmentally

    friendly, non-toxic, non-flammable, inexpensive, and

    readily available in high-purity from a number of

    sources. Product isolation is straightforward because

    CO2 is a gas under ambient conditions, removing the

    need for energy intensive drying steps. Moreover, addi-

    tional advantages are realized in the manufacture of flu-

    oropolymers such as increased safety in handling

    explosive monomers and enhanced polymer properties

    in many cases. As demonstrated CO2 is not only an ade-

    quate alternative, but in many cases a superior one. This

    Fig. 4. Emulsion-templated crosslinked polyacrylamide materials

    synthesized by polymerization of a high-internal phase CO2-in-water

    emulsion (C/W HIPE): (a) SEM image of sectioned material and (b)

    confocal image of same material, obtained by filling the pore structure

    with a solution of fluorescent dye. As such (a) shows the walls of the

    material while (b) show the holes formed by templating the sc-CO2emulsion droplets. Both images = 230lm 230 lm. Ratio of CO2/

    aqueous phase = 80:20 v/v.d Pore volume = 3.9 cm3/g. Average pore

    diameter = 3.9 lm. Adapted from Butler et al. [56].

    330 C.D. Wood et al. / Current Opinion in Solid State and Materials Science 8 (2004) 325331

  • 7/28/2019 [13]desimone

    7/7

    is demonstrated no more poignantly that DuPonts re-

    cent $275 million investment in its CO2 based TeflonTM

    production process. In addition, CO2 affords a number

    of specific advantages that can be derived from the use

    of SCFs for the synthesis and modification of porous

    materials. There are a number of specific chemical

    advantages that are realized by using CO2 for the syn-thesis of a range of materials but moreover, they offer

    the potential of reducing organic solvent usage in the

    production of materials.

    Acknowledgments

    We thank the NSF STC (grant number 537494), EPA

    (grant number 535325), ONR (grant number 535747),

    EPSRC. A.I.C. acknowledges the support of the Royal

    Society for the provision of a Research Fellowship.

    References

    [1] DeSimone JM. Science 2002;297:799803.

    [2] McHugh MA, Krukonis V. Supercritical fluid extraction. 2nd

    ed. Butterworth Heinemann; 1994.

    [3] Bartle KD, Clifford AA, Jafar SA, Shilstone GF. J Phys Chem,

    1991. Ref. Data 20713778.

    [4] Johnson KP, Penninger, editors. Supercritical fluid science and

    technology. Washington, DC: American Chemical Society; 1989.

    [5] Phelphs CL, Smart NG, Wai CM. J Chem Ed 1996;73:11638.

    [6] Kendall JL, Canelas DA, Young JL, DeSimone JM. Chem Rev

    1999;99:543.

    [7] DeYoung JP, Romack TJ, DeSimone JM. In: Hougham G,

    Cassidy PE, Johns K, Davidson T, editors. Fluoropolymers 1:synthesis. New York: Kluwer Academic/Plenum Publishers;

    1999. p. 191205.

    [8] Davidson TA, DeSimone JM. In: Jessop PG, Leitner W, editors.

    Chemical synthesis using supercritical fluids. New York, Chich-

    ester, Brisbane, Singapore, Toronto: Wiley-VCH Weinheim;

    1999. p. 297.

    [9] Cooper AI. Adv Mater 2003;15:1049.

    [10] Cooper AI. J Mater Chem 2000;10:207.

    [11] Wood CD, Yarbrough JC, Roberts GW, DeSimone JM. In:

    Kemmere M, Meyer T, editors. Polymer processes in carbon

    dioxide. Wiley: New York, NY; in press.

    [12] Resnick RP, Buck WH. In: Scheirs J, editor. Modern fluoropoly-

    mers. Chichester: John Wiley & Sons; 1997. p. 397.

    [13] Moody CA, Field JA. Environ Sci Technol 2000;34:386470.

    [14] Hogue CC & EN. August 30, 2004, p. 17.[15] DeYoung JP, Romack TJ, DeSimone JM. Polym Prepr (Am

    Chem Soc Div Polym Chem) 1997;38:424.

    [16] Fiering AE. In: Banks RE, Smart BE, Tatlow JC, editors.

    Organofluorine chemistry: principles and commercial applica-

    tions. New York: Plenum Press; 1994.

    [17] Van Bramer DJ, Schiflett MB, Yokozeki A. United States Patent

    5,345,013, 1994.

    [18] Romack TJ, DeSimone JM. Macromolecules 1995;28:8429.

    [19] Bunyard WC, Kadla JF, DeSimone JM. J Am Chem Soc

    2001;123:7199.

    [20] Wood CD, Michel U, Rolland JP, DeSimone JM. J Fluorine

    Chem 2004;125:1671.

    [21] DeSimone JM, Guan Z, Elsbernd C. Science 1992;257:945.

    [22] Romack TJ, DeSimone JM. Macromolecules 1995;28:842931.

    [23] Parkinson G. Chem Eng 1999;106:17.

    [24] Michel U, Resnick P, Kipp BE, DeSimone JM. Macromolecules

    2003;36:710713.

    [25] Charpentier PA, Kennedy KA, DeSimone JM, Roberts GW.

    Macromolecules 1999;32:59735.

    [26] Charpentier PA, DeSimone JM, Roberts GW. Ind Eng Chem Res

    2000;39:458896.

    [27] Saraf MK, Gerrard S, Wojcinski LM, Charpentier PA, DeSimone

    JM, Roberts GW. Macromolecules 2002;35:797685.

    [28] Guan Z, Elsbernd CS, DeSimone JM. Polym Prepr (Am Chem

    Soc, Div Polym Chem) 1992;34:329.

    [29] Guan Z, Combes JR, Elsbernd CS, DeSimone JM. Polym Prepr

    (Am Chem Soc Div Polym Chem) 1993;34:446.

    [30] Ehrlich P. Chemtracts: Org Chem 1993;6:92.

    [31] DeSimone JM. US Pat. 5,496,901, 1996.

    [32] DeSimone JM. US Pat. 5,688,879, 1997.

    [33] DeSimone JM. US Pat. 5,739,223, 1998.

    [34] McClain JB, Londono JD, Combes JR, Romack TJ, Canelas DA,

    Betts DE, et al. J Am Chem Soc 1996;118:917.

    [35] Rolland JP, Van Dam RM, Schorzman DA, Quake SR,

    DeSimone JM. J Am Chem Soc 2004;126:2322.

    [36] Rolland JP, Hagberg EC, Denison GM, Carter KR, DeSimone

    JM. Angew Chem Int Ed 2004;43:5796.

    [37] Sianesi D, Marchionni G, DePasquale RJ. In: Banks RE, Smart

    BE, Tatlow JC, editors. Organofluorine chemistry: principles and

    commercial applications. New York: Plenum Press; 1994. p. 431.

    [38] Bunyard WC, Romack TJ, DeSimone JM. Macromolecules

    1999;32:8224.

    [39] Cooper AI, Holmes AB. Adv Mater 1999;11:1270.

    [40] Cooper AI, Wood CD, Holmes AB. Ind Eng Chem Res

    2000;39:4741.

    [41] Svec F, Frechet JMJ. Science 1996;273:205.

    [42] Svec F, Frechet JMJ. Ind Eng Chem Res 1998;38:34.

    [43] Yu C, Xu MC, Svec F, Frechet JMJ. J Polym Sci A, Polym Chem

    2002;40:755.

    [44] Petro M, Svec F, Frechet JMJ. Biotechnol Bioeng 1996;49:355.

    [45] Hebb AK, Senoo K, Bhat R, Cooper AI. Chem Mater

    2003;15:2061.[46] Sherrington DC. Chem Commun 1998;21:2275.

    [47] Wood CD, Cooper AI. Macromolecules 2001;34:5.

    [48] Imhof A, Pine DJ. Adv Mater 1998;10:697.

    [49] Schmidt-Winkel P, Lukens WW, Yang PD, Margolese DI, Letlow

    JS, Ying JY, et al. Chem Mater 2000;12:686.

    [50] Manoharan VN, Imhof A, Thorne JD, Pine DJ. Adv Mater

    2001;13:447.

    [51] Zhang H, Hardy GC, Rosseinsky MJ, Cooper AI. Adv Mater

    2003;15:78.

    [52] Cameron NR, Sherrington DC. Adv Polym Sci 1996;126:163.

    [53] Busby W, Cameron NR, Jahoda CAB. Biomacromolecules

    2001;2:154.

    [54] Zhang H, Cooper AI. Chem Mater 2002;14:4017.

    [55] Lee CT, Psathas PA, Johnston KP, deGrazia J, Randolph TW.

    Langmuir 1999;15:6781.[56] Butler R, Davies CM, Cooper AI. Adv Mater 2001;13:1459.

    [57] Butler R, Hopkinson I, Cooper AI. J Am Chem Soc

    2003;125:14473.

    [58] Song J, Saiz E, Bertozzi CR. J Am Chem Soc 2003;125:1236.

    [59] Luo Y, Dalton PD, Shoichet MS. Chem Mater 2001;13:4087.

    [60] Chiellini F, Bizzari R, Ober CK, Schmaljohann D, Yu TY, Solaro

    R, et al. Macromol Rapid Commun 2001;22:1284.

    [61] Shaffer KA, Jones TA, Canelas DA, DeSimone JM. Macromol-

    ecules 1996;29:2704.

    [62] Sarbu T, Styranec T, Beckman EJ. Nature 2000;405:165.

    [63] Sarbu T, Styranec T, Beckman EJ. Ind Eng Chem Res

    2000;39:4678.

    [64] Bray CL, Tan B, Wood CD, Cooper AI. J Mater Chem

    2005;15:456.

    C.D. Wood et al. / Current Opinion in Solid State and Materials Science 8 (2004) 325331 331