Compatibility of medical-grade polymers with dense CO2

Compatibility of medical-grade polymers with dense CO2

Document
Talk
 
 
 
 
Year: 
Abstract: 

This study reports the effect of exposure to liquid carbon dioxide on the mechanical properties of selected medical polymers. The tensile strengths and moduli of fourteen polymers are reported. Materials were exposed to liquid CO2, or CO2 + trace amounts of aqueous H2O2, at 6.5 MPa and ambient temperature. Carbon dioxide uptake, swelling, and distortion were observed for the more amorphous polymers while polymers with higher crystallinity showed little effect from CO2 exposure. Changes in tensile strength were not statistically significant for most plastics, and most indicated good tolerance to liquid CO2. These results are relevant to evaluating the potential of liquid CO2-based sterilization technology.

DOI: 
10.1016/j.supflu.2007.05.002
Type of document: 
Language: 
J. of Supercritical Fluids 42 (2007) 366–372 Compatibility of medical-grade polymers with dense CO2 A. Jim´ nez a , G.L. Thompson a , M.A. Matthews a,∗ , T.A. Davis a , e K. Crocker b , J.S. Lyons b , A. Trapotsis c a b Department of Chemical Engineering, University of South Carolina, Columbia, SC 29208, USA Department of Mechanical Engineering, University of South Carolina, Columbia, SC 29208, USA c Consolidated Stills and Sterilizers, Boston, MA 02134, USA Received 31 July 2006; received in revised form 9 May 2007; accepted 10 May 2007 Abstract This study reports the effect of exposure to liquid carbon dioxide on the mechanical properties of selected medical polymers. The tensile strengths and moduli of fourteen polymers are reported. Materials were exposed to liquid CO2 , or CO2 + trace amounts of aqueous H2 O2 , at 6.5 MPa and ambient temperature. Carbon dioxide uptake, swelling, and distortion were observed for the more amorphous polymers while polymers with higher crystallinity showed little effect from CO2 exposure. Changes in tensile strength were not statistically significant for most plastics, and most indicated good tolerance to liquid CO2 . These results are relevant to evaluating the potential of liquid CO2 -based sterilization technology. © 2007 Elsevier B.V. All rights reserved. Keywords: Supercritical fluids; Carbon dioxide; Mechanical test; Polymethylmethacrylate; Polyvinylchloride; Polyethylene 1. Introduction Synthetic polymers are important in many biomedical applications, such as artificial joints and vascular implants. Medical devices used in invasive procedures must be cleaned and sterilized prior to use. Sterilization is the destruction of living organisms, and must be done without damaging the material surface and without compromising the bulk material strength or biocompatibility of implantable devices. It is well known that established sterilization processes may have adverse effects on certain biomaterials, including metals and polymers [1]. Common sterilization processes include steam autoclaving, gamma irradiation, and chemical treatments with ethylene oxide or hydrogen peroxide plasma [1–3]. Parameters such as temperature, pressure, energy intensity, sterilant concentration, and cycle time vary [4] and the process chemistry and conditions determine the applicability of each process for a given material or device. New sterilization processes continue to be investigated, with emphases on reducing the process temperature and minimizing contamination. A novel approach to sterilization is based on carbon dioxide (CO2 ) technology; reviews and patents have ∗ Corresponding author. Tel.: +1 803 777 0556; fax: +1 803 777 8265. E-mail address: matthews@engr.sc.edu (M.A. Matthews). 0896-8446/$ – see front matter © 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.supflu.2007.05.002 recently appeared [2,3,5–8]. The deactivation of vegetative bacteria using liquid CO2 in a single step has been demonstrated many times and a comprehensive review has been prepared by Zhang et al. [8]. Because of its potential for cleaning, disinfecting, or sterilizing at mild temperatures (typically between ambient and 79 ◦ C), and because of the cost advantages of liquid CO2 over supercritical, this study focuses on the bulk mechanical properties of selected polymeric biomaterials before and after exposure to liquid CO2 . CO2 technology is attractive in part because CO2 is non-flammable, non-toxic, physiologically safe, chemically inert and readily available [5,9–11]. When heated and compressed above its critical point (7.38 MPa and 304.2 K) CO2 exhibits a liquid-like density (0.6–1.0 × 10−3 kg m−3 ) [12] but gas-like diffusivity (10−7 to 10−8 m2 s−1 ) and viscosity (3–7 × 10−5 N s m−2 ), and zero surface tension [11]. These properties allow CO2 to penetrate porous structures easily. Typical process temperatures are on the order of 0–40 ◦ C, so there is the potential for developing a low-temperature sterilization technology. Research has shown that compressed CO2 kills many clinically relevant gram positive (Listeria monocytogenes, Staphylococcus aureus, and Enterococcus faecalis) [13,14] and gram negative (Salmonella typhimurium, E. coli, and Pseudomonas aeruginosa) vegetative bacteria [2,15]. Bacterial spores can also be sterilized with CO2 -based processes A. Jim´ nez et al. / J. of Supercritical Fluids 42 (2007) 366–372 e [2,7,16–18]. A 6-log reduction of B. pumilus, B. atrophaeus, and G. stearothermophilus spores was achieved at relatively low temperatures and treatment time (40 ◦ C, 27.58 MPa for 4 h) [19]. Several groups are now actively working on development of CO2 technology for biomaterial sterilization [7]. The objective of this work is to evaluate the tensile strength of several polymers before and after exposure to liquid CO2 , or to liquid CO2 plus aqueous H2 O2 . The polymers under investigation are used in manufacturing flexible surgical and diagnostic instruments, such as endoscopes. Loss of tensile strength will lead to degradation of the mechanical performance. The focus of the present study is on whether the materials will be tolerant to temperatures and elevated pressures (typically higher than 5.0 MPa) typical of liquid CO2 processing. Many polymers absorb significant amounts of gases and vapors when exposed at high pressures [11]. In particular, CO2 absorption can induce plasticization, swelling, and a decrease in the glass transition temperature [21]. Several reports examine the effect of high-pressure CO2 on certain specific polymers, including silicone rubber [20–23], cellulose acetate [24], poly(vinylidene fluoride) [20], poly(vinyl chloride) [20], poly(methyl methacrylate) [20,25], polystyrene [20,26], poly(vinyl benzoate) [27], low density polyethylene [28,29] and polysulfone [29,30]. The solubility of CO2 in polymers [24,29] and the resultant swelling and absorption behavior [23–25,29–32] have been recently investigated. 367 Fig. 1. Dimensions of specimens used for ASTM D638 tests. imens were prepared as specified by the ASTM D638 (Fig. 1). A high-pressure water jet was used to cut specimens; the use of the water jet allowed precision cutting at ambient temperature (i.e., without the friction heating from a conventional blade) and minimized the introduction of residual stresses in the samples. The specimens were washed in cool water to remove residue, dried, and weighed. The length, width (specimen maximum and minimum for dogbones) and thickness of the specimens were measured prior to CO2 processing. Also, specimens were weighed before treatment, immediately after treatment, and after degassing for 24 h post-treatment. Specimen thickness varied from 0.16 cm for UHMWPE to 0.688 cm for PET. Qualitative observations of color, unique markings and flexibility were also recorded. 2. Materials and methods 2.3. Processing with CO2 2.1. Materials A total of eighteen different commercial polymers were initially selected for study because of their relevance to the manufacture of flexible medical instruments. Four polymers were obtained from Laird Plastics (ultra-high molecular weigh polyethylene (UHMWPE), polystyrene (PS), low density polyethylene (LDPE), and high density polyethylene (HDPE)). The rest were obtained from Laboratory Devices Company (natural rubber latex (NRL), ethylene propylene diene monomer (EPDM) elastomer, poly(methyl methacrylate) (PMMA), silicone rubber (SI), polyurethane (PUR), acrylonitrile–butadiene–styrene (ABS) rubber, polycarbonate (PC), poly(ethylene terephthalate) (PET), poly(phenylene oxide) (PPO), poly(vinylidene fluoride) (PVDF), Teflon® , polyvinyl chloride (PVC), Nylon (PA), and acetal). Liquid CO2 from Holox (Coleman Grade, 99.99%) was used, and in some experiments aqueous hydrogen peroxide (H2 O2 , 30%) from Fisher Scientific was added. 2.2. ASTM D638 tensile testing ASTM Procedure D638 was used to evaluate tensile properties of polymers before and after exposure to liquid CO2 . Thickness was measured with a caliper (Central Tools, Incorporated) and mass was determined both before and after treatment with a Mettler AE 240 balance. Polymers were obtained in sheet form, from which the standard tensile “dogbone” spec- All specimens were exposed to liquid CO2 in a 1 L high-pressure vessel from Pressure Products Industries Inc., schematically represented in Fig. 2. The vessel is capable of operating in a range of 0–60 ◦ C (32–140 ◦ F) and up to 44.5 MPa (445 bar). Up to six specimens were mounted in the vessel for simultaneous exposure to the CO2 environment. Liquid CO2 from the CO2 supply was drawn through a 0.2 cm porous steel filter to a liquid CO2 pump (Thar Designs, P-200). The CO2 was preheated in 10 feet of coiled tubing to the desired temperature; a circulating bath (Fisher Scientific, Isotemp 1013S) provided temperature control. The pre-heated CO2 was filtered a second time with a 0.025 cm filter at the vessel inlet. Vessel temperature was maintained with an internal coil of tubing connected to the circulating bath. A thermocouple (Fluke 54) mounted in the built-in thermowell was used to monitor the internal temperature. An outlet at the top of the vessel was used to purge air before pressurization, and the drain valve at the bottom was used for depressurization. The temperature of the fluid exiting the drain valve was monitored by a thermocouple (Fluke 54) fastened into a high-pressure tee fitting. The vessel outlets and vessel bypass line were joined to allow flow through either a manual valve or a pressure regulator (Tescom, 26-1724-24-154) before venting to the atmosphere. The vessel was flushed with gaseous CO2 for about 10 s before pressurizing. The process temperature was ambient, typically between 17 and 24 ◦ C. To ensure that CO2 remained liquid, the pressure was set to 6.5 MPa. For experiments that required 368 A. Jim´ nez et al. / J. of Supercritical Fluids 42 (2007) 366–372 e Fig. 2. Schematic representation of the 1 L high-pressure reactor. H2 O2 , approximately 200 ␮L of aqueous H2 O2 (about 0.07 M) were injected at approximately 6 min into the pressurization step. To ensure uniform mixing of H2 O2 and CO2 , the internal sixblade impeller was rotated at about 525 rpm during treatments. The density and viscosity of CO2 at ambient temperature and 6.5 MPA are 0.797 g/mL and 1.5 × 10−5 kg/m s, respectively. Specimens were maintained at the prescribed conditions for 1 h after the pressure stabilized at 6.5 MPa, then the regulator-bypass valve was opened slowly to depressurize the vessel. The time required to depressurize the vessel (15–20 min) was recorded. The specimens were visually inspected immediately after opening the pressure vessel. Specimen weight and dimensions were recorded along with the time required for depressurization and removal from the vessel. Mechanical tests were initiated after a 24 h holding time to ensure that CO2 had desorbed. Any further changes in the specimen after 24 h were noted. Miscibility of aqueous H2 O2 and CO2 at these conditions was confirmed in separate experiments by direct observation in a Thar variable volume, stirred view cell shown in Fig. 3. CO2 was fed into the variable volume view cell with a magnetic mixer Fig. 3. Thar Technologies Inc. variable volume view cell. A. Jim´ nez et al. / J. of Supercritical Fluids 42 (2007) 366–372 e 369 using an ISCO 500HP syringe pump controlled by an ISCO series D controller. Ten microliters of 30% H2 O2 were carried quantitatively into the CO2 pressurized chamber through the use of a Valco Instruments (Houston, TX) six-port liquid injection valve. H2 O2 droplets observed at the beginning of the process disappeared once the process pressure was reached. 2.4. Mechanical testing Mechanical testing was performed on treated and control specimens. ASTM D638 tensile strength tests were performed with a 311 N-capacity Tinius Olsen universal testing machine, with a loading rate set at 0.005 m/min. Labtech NotebookTM was used for data acquisition. At least four replicate specimens were used for the untreated controls and for each CO2 -treated material. An extensometer with a 10% strain range was used to record the strain during the initial elastic response of each specimen. This strain and force data were used to calculate the modulus of elasticity. The extensometer was removed prior to sample yielding (i.e., at approximately 9% strain) and each specimen was further loaded until failure. The peak force was recorded and used to calculate the ultimate tensile strength. The crosssectional dimensions at the point of rupture were measured to determine the percent-reduction in area after fracture. As will be described below, four of the polymers proved incompatible with CO2 processing and therefore were not subjected to a full range of mechanical testing. 2.5. Data analysis Quantitative data presented in this work are the results of at least four replicate tests. As indicated for the specific results shown in the figures, certain means of different treatments were compared using a two-sided Student’s t-test with a confidence level α = 0.01. All the results are presented as mean ± S.D. 3. Results and discussion The results of all tests are summarized in the form of bar graphs in Figs. 4 and 6–8. All bars represent the average property of at least four replicate tests, and in many cases if sufficient material was available, five or six specimens were tested. The error bars shown represent one standard deviation. Fig. 4 shows the weight change after treatment with CO2 or CO2 /H2 O2 followed by degassing for 24 h. The “X” in Fig. 4 designates a statistically significant difference between treatments (CO2 , CO2 + H2 O2 ), as determined by a two-sided Student’s t-test with a confidence level α = 0.01. The more amorphous polymers (Latex rubber, PMMA, ABS, PUR, EPDM, PS, PC, LDPE, PPO, PTFE and PVC) are on the left, and the more crystalline polymers (UHMWPE, HDPE, POM, PVDF, PET, PA, SI) are to the right in Fig. 4. Amorphous polymers have little molecular orientation and large free volume, thus CO2 penetrates even if no specific intermolecular interactions are present. Crystalline polymers, in contrast, have highly ordered molecular arrangement and relatively less free volume, so CO2 will not absorb as readily [33]. Fig. 4 reflects these generalizations. Fig. 4. Weight change due to uptake of CO2 in the specimens. The “X” designates a statistically significant difference between treatments. Latex rubber gained more mass than any other polymer, about 7% under pure CO2 and about 5.5% under CO2 /H2 O2 . The behavior of PMMA was comparable to that for latex rubber; PMMA showed a mass uptake of over 6% when exposed to either CO2 or to CO2 plus H2 O2 . PMMA was slightly distorted and softened, and specimens were deformed at areas of contact with the reactor. ABS, polyurethane and EPDM also gained significant mass, on the order of 5%. EPDM deformed during treatment, but polyurethane and ABS did not deform. Of the amorphous polymers, latex rubber specimens were most swollen and distorted. After 2 days, the specimens returned to their original sizes and shapes. Because of excessive swelling, latex rubber was not subjected to tensile tests. With crystalline polymers, the weight uptake was small and there were no observable changes in the materials after CO2 or CO2 plus H2 O2 treatment. The ability of a polymer to solubilize CO2 depends on both chemical interactions and the degree of crystallinity. Polymers possessing electron donating or Lewis base functional groups, for example, exhibit specific attractive interactions with CO2 [34]. The weight gain in latex rubber, PMMA and EPDM can be interpreted in terms of intermolecular forces. Latex rubber and EPDM are non-polar polymers and do not exhibit any specific intermolecular interaction with CO2 . The weight increase after exposure of CO2 is attributed primarily to the amorphous structure. The degree of cross-linking and effects of chain branching also increase the free volume. PMMA contains a carbonyl functional group in the backbone chain, which is an electron-donor or Lewis base. The presence of this carbonyl group enhances the solubility of CO2 and induces a change in weight after treatment. Compared to crystalline polymers, amorphous polymers show more weight gain and swelling due to CO2 absorption. The more crystalline polymers, nylon and PET, showed the least weight change. Polar amide groups in the backbone chain of nylon form strong hydrogen bonds that promote crystallinity. In PET, crystallinity is induced by the polar ester groups as well as the orderly arrangement of aromatic rings. Highly crystalline lattices inhibit penetration and adsorption of CO2 . A practical sterilization process should be accomplished in as short a time as possible; significant CO2 uptake and degassing times are detrimental. However, previous work on sterilization processes [8,35] and mechanisms [19] indicate that processing 370 A. Jim´ nez et al. / J. of Supercritical Fluids 42 (2007) 366–372 e times on the order of 0 s to 4 h may be necessary to achieve complete sterilization. It is therefore of interest to estimate the relative degree of CO2 absorption under the experimental conditions used in this work. In a simple model, the specimens are treated as flat planes, and Fickian diffusion is assumed. The Fickian sorption and desorption of fluids into or out of a flat polymeric membrane can be modeled according to Crank and Park [36] by the following equation: Mt 8 =1− 2 M∞ π ∞ e−D((2n+1/2)π/L) (2n + 1)2 n=0 2t (1) In this equation Mt is the total amount of diffusing fluid that has entered the polymeric membrane at time t, M∞ represents the saturation amount of CO2 that can be absorbed, L the sample thickness and D is the diffusion coefficient. In this dimensionless model, Mt /M∞ describes the extent of CO2 adsorption at a time t relative to the maximum equilibrium absorption at t = ∞. The important parameter in this model is D. The diffusion coefficient of SC CO2 has been measured in a variety of polymers, including PVC with a diffusion coefficient of 0.24 × 10−11 m2 s−1 [37]. As a sensitivity check, Eq. (1) is also solved using D = 0.24 × 10−9 m2 s−1 ; this higher diffusion coefficient is representative of CO2 diffusing in an amorphous polymer. Fig. 5 shows calculations of Mt /M∞ over time for PVC for L = 0.0016 m, i.e., the thinnest specimen tested. As observed from this figure, after 1 h the amount of CO2 adsorbed is approximately 10% of the equilibrium amount. Since PVC is crystalline, CO2 will not easily penetrate. As observed in Fig. 5, the treatment time required for the CO2 to completely penetrate the crystalline matrix is about a month. In contrast, CO2 uptake is comparatively rapid in the case of D = 0.24 × 10−9 m2 s−1 . For amorphous polymers, the amount of CO2 adsorbed is about 70% of the equilibrium amount after 1 h and the treatment time necessary for the CO2 to completely penetrate the matrix is about a day. This behavior is consistent with the swelling of polymers shown in Fig. 4. It is emphasized that this is a simple model for order of magnitude calculations. Fig. 6 shows the ASTM D638 tensile strength of fourteen plastics (latex, PUR, and SI are not reported because they were Fig. 5. CO2 uptake in a planar polymer as predicted by the Crank and Park [36] model. The diffusion coefficient of SC CO2 in PVC is 0.24 × 10−11 m2 s−1 ; D = 0.24 × 10−9 m2 s−1 is representative of CO2 diffusing in an amorphous polymer. Fig. 6. Tensile strength of plastic specimens, as measured by ASTM D638. The “X” designates a statistically significant change compared to the control. too elastic, while EPDM was too brittle). Fig. 6 shows data for CO2 - and CO2 plus H2 O2 -treated specimens, respectively. An “X” above a bar indicates a statistically significant difference from the untreated control group, as determined by a two-sided Student’s t-test with a confidence level (α) of 0.01. Each bar represents the average of 4–6 replicate specimens. The tensile strength of PMMA decreased about 19% after exposure to CO2 , while for ABS it decreased about 10%. The tensile strength of PC shows a statistically significant decrease of about 6% after exposure to either treatment. PVC shows a statistically significant decrease, similar to PC. PET shows a very slight decrease in tensile strength when treated with CO2 plus H2 O2 . Other plastics did not show statistically significant changes in tensile strength. Fig. 7 shows the area reduction of the specimens after treatment. This is the decrease in the cross-sectional area of the neck of the dogbone (Fig. 1) after failure. Rigid polymers like PMMA and PS showed only slight area reduction before failure. The area reduction increases with the pliability of the material, and so the strain before failure will likewise increase. Note that PMMA and PS, for example, show very little area reduction. CO2 plus H2 O2 treatment increased the area reduction in PC and HDPE. ABS and HDPE, after treatment with pure CO2 , show a statistically significant (though small numerically) increase in the area reduction. The area reduction for the remaining polymers was not affected by CO2 processing. Fig. 7. Area reduction in plastic specimens. The “X” designates a statistically significant change compared to the control. A. Jim´ nez et al. / J. of Supercritical Fluids 42 (2007) 366–372 e 371 containing those polymers. It should be emphasized that all these data are for a single cycle of exposure to liquid CO2 . For development of a process to sterilize reusable instruments, it remains to be demonstrated that a material or instrument could withstand multiple cycles. Finally, it should be noted that liquid CO2 may extract low molecular weight oligomers, plasticizers, or other components blended with commercial polymers. There is evidence that extraction occurred with certain systems as indicated by the weight changes shown in Fig. 4, namely polystyrene, polyurethane, acrylic, silicone, latex, LDPE, HDPE, PET, PPO, acetal, PVDF, and EPDM. Acknowledgments Fig. 8. Modulus of elasticity of plastic specimens. The “X” designates a statistically significant change compared to the control. Fig. 8 shows Young’s modulus (modulus of elasticity) calculated from the stress/strain curve. The modulus is strongly correlated with variations in the polymer structure, the molecular weight and the density among specimens of the same material [38]. PMMA showed a statistically significant decrease in the modulus after exposure to CO2 , suggesting CO2 plasticization [30]. PS demonstrated a statistically significant increase in the modulus when treated with CO2 plus H2 O2 . PVC also showed a statistically significant increase in modulus after exposure to pure CO2 . 4. Conclusions The mechanical properties of several medically relevant polymers have been evaluated before and after exposure to liquid CO2 at ambient temperature and 6.5 MPa. There was appreciable CO2 uptake even after 24 h of degassing for some of the amorphous polymers (latex rubber, ABS, PMMA, polyurethane, and EPDM). Crystalline polymers showed little weight change. A statistically significant decrease in the tensile strength was observed for PMMA (CO2 ), ABS (CO2 /H2 O2 ), PC (CO2 /H2 O2 ), PVC (CO2 /H2 O2 ), and this may be clinically significant also. PET (CO2 /H2 O2 ) showed a slight decrease in tensile strength. PMMA showed a statistically significant decrease in Young’s modulus after exposure to CO2 , probably because of induced plasticization, whereas PS (CO2 /H2 O2 ) and PVC (CO2 ) demonstrated a statistically significant increase in modulus when treated with CO2 plus H2 O2 that could be due to an increase in the degree of crystallinity. Liquid CO2 , proposed for cleaning and sterilizing medical devices, may be used at realistic processing conditions with certain polymeric biomaterials. However, one must take into account the interactions between CO2 and the materials of interest. Crystalline polymers show little if any uptake, swelling and distortion, and devices made from these materials would likely be compatible with a CO2 -based process. In contrast, some amorphous polymers showed more pronounced effects with CO2 that could compromise bulk material strength, and a CO2 sterilization process might not be feasible for medical devices This work was supported in part by Consolidated Stills and Sterilizers and by the NIH/National Institute of Biomedical Imaging and Bioengineering under a Bioengineering Research Partnership grant (R01EB55201). Financial assistance for Aidaris Jim´ nez was provided by the South Carolina e EPSCOR/BRIN program. References [1] Y.H. An, F.I. Alvin, Q.K. Kang, M. LaBerge, M.J. Drews, J. Zhang, M.A. Matthews, C.R. Arciola, Effects of sterilization on implant mechanical property and biocompatibility—a concise review, Int. J. Artif. Organs 28 (2005) 1126–1137. [2] A.K. Dillow, F. Dehghani, J.S. Hrkach, N.R. Foster, R. Langer, Bacterial inactivation by using near and supercritical carbon dioxide, Proc. Natl. Acad. Sci. U.S.A. 96 (1999) 10344–10348. [3] M.A. Matthews, L.S. Warner, H. Kaiser, Exploring the feasibility of using dense-phase carbon dioxide for sterilization, Med. Device Diagn. Ind. 5 (2001) 140–149. [4] L.A. Feldman, H.K. Hui, Compatibility of medical devices and materials with low-temperature hydrogen peroxide gas plasma, Med. Device Diagn. Ind. Mag. (1997) 57–70. [5] N. Elvassore, S. Sartorello, S. Spilimbergo, A. Bertucco, Micro-organisms inactivation by supercritical CO2 in a semi-continuous process, in: M. Perrut, E. Reverchon (Eds.), Proceedings of the Seventh Meeting on Supercritical Fluids, Antibes, France, 2000, pp. 773–778. [6] A. Schmidt, K. Beermann, E. Bach, E. Schollmeyer, Disinfection of textile materials contaminated with E. coli in liquid carbon dioxide, J. Cleaner Prod. 13 (2005) 881. [7] S. Spilimbergo, A. Bertucco, Non-thermal bacterial inactivation with dense CO2 , Biotechnol. Bioeng. 84 (2003) 627–638. [8] J. Zhang, T.A. Davis, M.A. Matthews, M.J. Drews, M. LaBerge, Y.H. An, Sterilization using high-pressure carbon dioxide, J. Supercrit. Fluids 38 (2006) 354–372. [9] A. Enomoto, K. Nakamura, K. Nagai, T. Hashimoto, M. Hakoda, Inactivation of food microorganisms by high-pressure carbon dioxide treatment with or without explosive decompression, Biosci. Biotechnol. Biochem. 61 (1997) 1133–1137. [10] S.I. Hong, W.S. Park, Y.R. Pyun, Non-thermal inactivation of Lactobacillus plantarum as influenced by pressure and temperature of pressurized carbon dioxide, Int. J. Food Sci. Technol. 34 (1999) 125–130. [11] M. McHugh, V. Krukonis, Supercritical Fluid Extraction, 2nd ed., Butterworth-Heinemann, Newton, MA, 1994, pp. 1–16. [12] R. Span, W. Wagner, A new equation of state for carbon dioxide covering the fluid region from the triple-point temperature to 1100 K at pressures up to 800 MPa, J. Phys. Chem. Ref. Data 25 (1996) 1509– 1596. [13] O. Erkmen, Effect of carbon dioxide pressure on Listeria monocytogenes in physiological saline and foods, Food Microbiol. 17 (2000) 589–596. 372 A. Jim´ nez et al. / J. of Supercritical Fluids 42 (2007) 366–372 e [14] O. Erkmen, Antimicrobial effect of pressurized carbon dioxide on Staphylococcus aureus in broth and milk, Food Sci. Technol. (Lond.) 30 (1997) 826–829. [15] O. Erkmen, Inactivation of Salmonella typhimurium by high-pressure carbon dioxide, Food Microbiol. 17 (2000) 225–232. [16] A. Driks, Bacillus subtilis spore coat, Microbiol. Mol. Biol. Rev. 63 (1999) 1–20. [17] H. Ishikawa, M. Shimoda, K. Tamaya, A. Yonekura, T. Kawano, Y. Osajima, Inactivation of Bacillus spores by the supercritical carbon dioxide micro-bubble method, Biosci. Biotechnol Biochem. 61 (1997) 1022– 1023. [18] M. Kamihira, M. Taniguchi, T. Kobayashi, Sterilization of microorganisms with supercritical carbon dioxide, Agric. Biol. Chem. 51 (1987) 407– 412. [19] J. Zhang, N. Dalal, C. Gleason, M.A. Matthews, L.N. Waller, K. Fox, A. Fox, M.J. Drews, M. LaBerge, Y.H. An., On the mechanisms of deactivation of Bacillus atrophaeus spores using supercritical carbon dioxide, J. Supercrit. Fluids (2006) 268–273. [20] J.S. Chiou, J.W. Barlow, D.R. Paul, Plasticization of glassy polymers by carbon dioxide, J. Appl. Polym. Sci. 30 (1985) 2633–2642. [21] G.K. Fleming, W.J. Koros, Dilation of polymers by sorption of carbon dioxide at elevated pressures. 1. Silicone rubber and unconditioned polycarbonate, Macromolecules 19 (1986) 2285–2291. [22] M. Tang, T.-B. Du, Y.-P. Chen, Sorption and diffusion of supercritical carbon dioxide in polycarbonate, J. Supercrit. Fluids 28 (2004) 207–218. [23] R.G. Wissinger, M.E. Paulaitis, Swelling and sorption in polymer–carbon dioxide mixtures at elevated pressures, J. Polym. Sci., Part B: Polym. Phys. 25 (1987) 2497–2510. [24] S.A. Stern, A.H. De Meringo, Solubility of carbon dioxide in cellulose acetate at elevated pressures, J. Polym. Sci. Polym. Phys. Ed. 16 (1978) 735–751. [25] W.J. Koros, G.N. Smith, V. Stannett, High-pressure sorption of carbon dioxide in solvent-cast poly(methyl methacrylate) and poly(ethyl methacrylate) films, J. Appl. Polym. Sci. 26 (1981) 159–170. [26] W.C.V. Wang, E.J. Kramer, W.H. Sachse, Effect of high-pressure carbon dioxide on the glass transition temperature and mechanical properties of polystyrene, J. Polym. Sci. Polym. Phys. Ed. 20 (1982) 1371–1384. [27] Y. Kamiya, K. Mizoguchi, Y. Naito, T. Hirose, Gas sorption in poly(vinyl benzoate), J. Polym. Sci., Part B: Polym. Phys. 24 (1986) 535–547. [28] T. Hirose, K. Mizoguchi, Y. Kamiya, Dilation of polyethylene by sorption of carbon dioxide, J. Polym. Sci., Part B: Polym. Phys. 24 (1986) 2107–2115. [29] Y. Kamiya, T. Hirose, K. Mizoguchi, Y. Naito, Gravimetric study of highpressure sorption of gases in polymers, J. Polym. Sci., Part B: Polym. Phys. 24 (1986) 1525–1539. [30] J.S. Chiou, J.W. Barlow, D.R. Paul, Polymer crystallization induced by sorption of carbon dioxide gas, J. Appl. Polym. Sci. 30 (1985) 3911–3924. [31] A.R. Berens, G.S. Huvard, R.W. Korsmeyer, F.W. Kunig, Application of compressed carbon dioxide in the incorporation of additives into polymers, J. Appl. Polym. Sci. 46 (1992) 231–242. [32] J.H. Hildebrand, J.M. Prausnitz, R.L. Scott, Regular Related Solutions, Van Nostrand Reinhold Co., New York, 1970. [33] S. Doroudiani, C.B. Park, M.T. Kortschot, Effect of the crystallinity and morphology on the microcellular foam structure of semicrystalline polymers, Polym. Eng. Sci. 36 (1996) 2645–2662. [34] S.G. Kazarian, M.F. Vincent, F.V. Bright, C.L. Liotta, C.A. Eckert, Specific intermolecular interaction of carbon dioxide with polymers, J. Am. Chem. Soc. 118 (1996) 1729–1736. [35] J. Zhang, S. Burrows, C. Gleason, M.A. Matthews, M.J. Drews, M. LaBerge, Y.H. An., Sterilizing Bacillus pumilus spores using supercritical carbon dioxide, J. Microbiol. Methods (2006) 479–485. [36] J. Crank, G.S. Park (Eds.), Diffusion in Polymers, Academic Press, London, 1968, p. 452. [37] O. Muth, T. Hirth, H. Vogel, Investigation of sorption and diffusion of supercritical carbon dioxide into poly(vinyl chloride), J. Supercrit. Fluids 19 (2001) 299–306. [38] S.I. Seok, C.H. Ahn, M.Y. Jin, C.J. Lee, Y. Kang, Effect of molecular weight on the mechanical properties of MSSQ films, Mater. Chem. Phys. 84 (2004) 259–262.
Coments go here: