Novel Supercritical Carbon Dioxide Impregnation Technique for the

Analysis was carried out using PE Pyris Thermal Analysis software, version 10.1, ... 0.2 °C. A formulation mass (180–355 μm size fraction) equival...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIV OF PITTSBURGH

Article

A Novel Supercritical Carbon Dioxide Impregnation Technique for the Production of Amorphous Solid Drug Dispersions: A Comparison to Hot Melt Extrusion. Catherine Potter, Yiwei Tian, Gavin Walker, Colin McCoy, Peter Hornsby, Conor Donnelly, David S. Jones, and Gavin P. Andrews Mol. Pharmaceutics, Just Accepted Manuscript • DOI: 10.1021/mp500644h • Publication Date (Web): 02 Mar 2015 Downloaded from http://pubs.acs.org on March 11, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Molecular Pharmaceutics is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

A Novel Supercritical Carbon Dioxide Impregnation Technique for the Production of Amorphous Solid Drug Dispersions: A Comparison to Hot Melt Extrusion.

Catherine Potterab, Yiwei Tiana, Gavin Walkerc, Colin McCoya Peter Hornsbyb, Conor Donnellya, David S. Jonesa, and Gavin P. Andrewsa* a

The Drug Delivery and Biomaterials Group, School of Pharmacy, Medical Biology Centre,

Queen’s University, 97 Lisburn Road, Belfast. BT9 7BL, Northern Ireland, UK. b

School of Mechanical and Aeronautical Engineering, Queen’s University, 97 Lisburn Road,

Belfast. BT9 7BL, Northern Ireland, UK c. Department of Chemical and Environmental Science, University of Limerick, Ireland.

*Correspondence to: Gavin P. Andrews Tel: +44 (0) 28 9097 2646; Fax: +44 (0) 28 9024 7794; E-mail: [email protected]

Keywords: Supercritical Fluid Impregnation, Hot Melt Extrusion, Solid Dispersions, Soluplus, PVP.

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract The formulation of BCS Class II drugs as amorphous solid dispersions has been shown to provide advantages in respect of improving the aqueous solubility of these compounds. While hot melt extrusion (HME) and spray drying (SD) are among the most common methods for the production of amorphous solid dispersions (ASDs), the high temperatures often required for HME can restrict the processing of thermally labile drugs, whilst the use of toxic organic solvents during SD can impact on end-product toxicity. In this study, we investigated the potential of supercritical fluid impregnation (SFI) using carbon dioxide as an alternative process for ASD production of a model poorly water-soluble drug, Indomethacin (INM). In doing so, we produced ASDs without the use of organic solvents and at temperatures considerably lower than those required for HME. Previous studies have concentrated on the characterisation of ASDs produced using HME or SFI but have not considered both processes together. Dispersions were manufactured using two different polymers, Soluplus® and Polyvinylpyrrolidone K15 using both SFI and HME and characterised for drug morphology, homogeneity, presence of drug-polymer interactions, glass transition temperature, amorphous stability of the drug within the formulation and nonsink drug release to measure the ability of each formulation to create a supersaturated drug solution. Fully amorphous dispersions were successfully produced at 50% w/w drug loading using HME and 30% w/w drug loading using SFI. For both polymers formulations containing 50% w/w INM, manufactured via SFI, contained drug in the γ-crystalline form. Interestingly, there were lower levels of crystallinity in PVP dispersions relative to SOL. FTIR was used to probe for the presence of drug-polymer interactions within both polymer systems. For PVP systems, the nature of these interactions depended upon processing method, however for Soluplus formulations this was not the case. The area under the dissolution curve (AUC) was used as a measure of the time during which a supersaturated concentration could be maintained and for all systems SFI formulations performed better than similar HME formulations.

ACS Paragon Plus Environment

Page 2 of 41

Page 3 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Introduction The low aqueous solubility of many new chemical entities is a major issue currently being faced by the pharmaceutical industry with recent estimates suggesting that 90% of drugs within pharmaceutical pipelines possess limited aqueous solubility. It is also approximated that 40% of marketed oral drug products are considered as practically insoluble ( SOL HME. The disparity between supersaturation concentrations reached for the two polymers is attributed to the greater hydrophilicity of PVP over Soluplus and reflects similar results from Guo et al. (2013) who observed a slower carrier-controlled dissolution from a DiflusinalSoluplus formulation than from Diflusinal-PVP [32]. The drug release curves observed in this work (Figure 10) do not exhibit the spring and parachute shape and instead display a continuous increase in drug concentration over the full time of dissolution. It was also observed that none of the formulations dissolved completely over the course of the dissolution. It is assumed that neither the drug nor the polymer dissolve completely over eight hours in acidic conditions. It has been observed that the shape of the dissolution curve and drug concentration measured can be highly dependent on the method of measuring the dissolved drug [44]. Alonzo et al. (2011) observed very different dissolution profiles by first using an in situ UV-

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Vis method and subsequently by encasing the UV probe in a dialysis membrane. The much higher apparent readings gathered without the dialysis membrane in place were attributed to interference from drug molecules contained in polymeric nanoparticles rather than those truly dissolved in the dissolution medium. With the dialysis membrane in place the initial rate and extent of drug dissolution measured was greatly reduced and the dissolution behaviour was broadly linear. A number of the dissolution profiles gained in this work were also linear so to provide a more quantitative analysis linear regression analysis was carried out. The dissolution profiles of all formulations showed a good correlation to a linear fit except for 10 PVP SFI and 50 PVP SFI (Table 7) which exhibited an initial burst release. While all PVP formulations attained similar final concentrations after 8 hours, the higher initial release rate meant that these were also the formulations with the greatest AUC8hr value. The higher rate of release of drug from the SFI formulations was attributed to surface area differences which enabled faster dissolution of the polymer and is in agreement with the typically porous nature of supercritical fluid processed products [20]. It also contributes to the hypothesis that the dissolution from these formulations was carrier controlled. The 50 SFI formulations, which contained some crystalline material showed better dissolution than some of the completely amorphous lower drug loadings. This is in agreement with Pina et al. (2014) who showed that for formulations with similar drug but different crystalline content, the completely amorphous formulation did not always show the best dissolution [50] with dissolution performance being attributed to the polymer rather than the morphology of the API.

Amorphous stability studies The ten completely amorphous formulations were subjected to conditions of 40 ⁰C and 75 %RH for four months, after which the presence of crystalline content was investigated using HyperDSC. Raman spectroscopy was not a suitable characterisation technique for this study as the annealed samples exhibited Raman fluorescence. HyperDSC traces for the Soluplus formulations are illustrated in Figure 12. A Tg is apparent around 30-90 ⁰C and, if present, the INM endotherm appears in the area bounded by dotted lines. The enthalpy associated with the endotherms gained for each formulation are summarised in Table 8. 10 SOL SFI remained completely amorphous while the average crystalline content found in 10 SOL HME measured an enthalpy 0.09 J/g. One of the 10 SOL HME replicates was completely amorphous so crystallinity was not uniformly dispersed throughout the

ACS Paragon Plus Environment

Page 18 of 41

Page 19 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

formulation. All other formulations showed some degree of crystallinity. Considering the 10 and 30% w/w loaded formulations, a two way ANOVA showed statistically significant differences between drug loading but not between processing methods. Therefore it can be concluded that Soluplus formulations produced using SFI show a similar stability to those produced using hot melt extrusion. This is in agreement with the parity of strength of interactions indicated by FTIR in both SOL HME and SOL SFI formulations. A two way ANOVA confirmed statistical differences in the means of melting enthalpies of PVP formulations on both a drug loading and processing method basis. As indicated in Table 8, a larger endotherm was observed for SFI formulations in comparison to similar HME formulations. A difference in processing methods was not shown by Soluplus formulations and this disparity in behaviour is attributed to the hydrophilicity of PVP. PVP is an extremely hydrophilic powder and becomes tacky after exposure to room temperature and humidity. Formulations were pressed into a disk prior to the stability study and while the Soluplus formulations maintained their structural integrity, similar PVP formulations became more fluid and flowed to take the shape of their container. The relative instability is therefore attributed to the greater mobility induced by the uptake of water by the more hydrophilic polymer rather than to thermodynamic factors. The greater degree of crystallisation which occurred in PVP SFI formulations in particular is attributed to the greater surface area of SFI formulations which initially allowed a greater rate of intake of moisture. It is also in agreement with FTIR studies which suggested that relatively stronger interactions were generated through processing using HME than using SFI. Sinclair et al. (2011) [52] acknowledge the significant effect of increasing humidity on the amorphous stability of an amorphous solid dispersion of ibipinabant formulated using PVP, noting that moisture had a significant impact on the amorphous stability of the API.

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Conclusion Amorphous solid dispersions of INM with PVP and Soluplus have been successfully produced using HME and SFI. While treatment of γ-INM with scCO2 did not produce the amorphous form of the API, a morphology change was observed on the introduction of polymeric excipients Soluplus and PVP into the reaction chamber. While relatively higher amorphous drug loadings could be incorporated using HME the temperatures required were up to 100 ⁰C higher than for SFI. For INM-PVP the increased temperatures appeared to generate greater interaction between the species with FTIR showing more hydrogen bonding for HME formulations than SFI. Of the completely amorphous formulations a greater homogeneity was shown by extruded formulations over those produced using supercritical fluids highlighting the relative difficulty associated with processing using supercritical fluids and high viscosity polymers without the use of surfactants. PVP formulations showed a higher dissolution rate and final concentration than Soluplus formulations, the former producing a 4-4.5 times enhancement of crystalline solubility in comparison to two times for Soluplus. The difference between the two sets of carrierdependent dissolutions was attributed to the difference in dissolution rate of each polymer, PVP being more hydrophilic and being of a smaller molecular weight. Using Soluplus as the stabilising excipient, no difference in processing method was found for amorphous stability, SFI formulations exhibiting similar recrystallisation rates to extruded formulations. HME is useful method of production of ASDs due to its continuous, high throughput nature and the increasing range of equipment available allows for a greater range of APIs and excipients to be processed successfully. However scCO2 impregnation is a promising production method for thermally labile compounds or high glass transition polymers that would traditionally require high levels of plasticiser. It is useful to consider a range of processing methods for formulation of poorly aqueous soluble drugs as not all methods will be suitable for all formulations.

ACS Paragon Plus Environment

Page 20 of 41

Page 21 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

References 1. T. Loftsson, M.E. Brewster, Pharmaceutical applications of cyclodextrins: basic science and product development, J. Pharm. Pharmacol. 62 (2010) 1607-1621. 2. Y. Tian, J. Booth, E. Meehan, D.S. Jones, S. Li, G.P. Andrews, Construction of DrugPolymer Thermodynamic Phase Diagrams Using Flory-Huggins Interaction Theory: Identifying the Relevance of Temperature and Drug Weight Fraction to Phase Separation within Solid Dispersions, Molecular Pharmaceutics 10 (2013) 236-248. 3. Y. Tian, V. Caron, D.S. Jones, A. Healy, G.P. Andrews, Using Flory-Huggins phase diagrams as a pre-formulation tool for the production of amorphous solid dispersions: a comparison between hot-melt extrusion and spray drying, J. Pharm. Pharmacol. 66 (2014) 256-274. 4. M.M. Crowley, F. Zhang, M.A. Repka, S. Thumma, S.B. Upadhye, S.K. Battu, J.W. McGinity, C. Martin, Pharmaceutical applications of hot-melt extrusion: Part I, Drug Dev. Ind. Pharm. 33 (2007) 909-926. 5. D. Ma, A. Djemai, C.M. Gendron, H. Xi, M. Smith, J. Kogan, L. Li, Development of a HPMC-based controlled release formulation with hot melt extrusion (HME), Drug Dev. Ind. Pharm. 39 (2013) 1070-1083. 6. International Conference on Harmonisation, Geneva, ICH Impurities: Guideline for residual solvents Q3C (R5) Current Step 4 version dated 4 February 2011 (2011). 7. D. Mou, H. Chen, J. Wan, H. Xu, X. Yang, Potent dried drug nanosuspensions for oral bioavailability enhancement of poorly soluble drugs with pH-dependent solubility, Int. J. Pharm. 413 (2011) 237-244. 8. V. Caron, L. Tajber, O.I. Corrigan, A.M. Healy, A Comparison of Spray Drying and Milling in the Production of Amorphous Dispersions of Sulfathiazole/Polyvinylpyrrolidone and Sulfadimidine/Polyvinylpyrrolidone, Molecular Pharmaceutics 8 (2011) 532-542. 9. R. Vehring, Pharmaceutical particle engineering via spray drying, Pharm. Res. 25 (2008) 999-1022. 10. S. Kazarian, N. Brantley, B. West, M. Vincent, C. Eckert, In situ spectroscopy of polymers subjected to supercritical CO2: Plasticization and dye impregnation, Appl. Spectrosc. 51 (1997) 491-494. 11. S. Santoyo, L. Jaime, M.R. Garcia-Risco, A. Ruiz-Rodriguez, G. Reglero, Antiviral Properties of Supercritical Co2 Extracts from Oregano and Sage, Int. J. Food Prop. 17 (2014) 1150-1161. 12. K.M. Sharif, M.M. Rahman, J. Azmir, A. Mohamed, M.H.A. Jahurul, F. Sahena, I.S.M. Zaidul, Experimental design of supercritical fluid extraction - A review, J. Food Eng. 124 (2014) 105-116. 13. A.M.V. Schou-Pedersen, J. Ostergaard, M. Johansson, S. Dubant, R.B. Frederiksen, S.H. Hansen, Evaluation of supercritical fluid chromatography for testing of PEG adducts in pharmaceuticals, J. Pharm. Biomed. Anal. 88 (2014) 256-261.

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

14. Y. Reinwald, R.K. Johal, A.M. Ghaemmaghami, F.R.A.J. Rose, S.M. Howdle, K.M. Shakesheff, Interconnectivity and permeability of supercritical fluid-foamed scaffolds and the effect of their structural properties on cell distribution, Polymer 55 (2014) 435-444. 15. D.W. Pu, M.P. Devitt, S.C. Thickett, F.P. Lucien, P.B. Zetterlund, Dispersion polymerization of styrene in CO2-expanded ethanol, Polymer 54 (2013) 6689-6694. 16. S. Chang, T. Hsu, Y. Chu, H. Lin, M. Lee, Micronization of aztreonam with supercritical anti-solvent process, Journal of the Taiwan Institute of Chemical Engineers 43 (2012) 790797. 17. N. Esfandiari, S.M. Ghoreishi, Kinetic Modeling of the Gas Antisolvent Process for Synthesis of 5-Fluorouracil Nanoparticles, Chem. Eng. Technol. 37 (2014) 73-80. 18. J. Yu, Y. Guan, S. Yao, Z. Zhu, Preparation of Roxithromycin-Loaded Poly(l-lactic Acid) Films with Supercritical Solution Impregnation, Ind Eng Chem Res 50 (2011) 13813-13818. 19. M. Banchero, L. Manna, S. Ronchetti, P. Campanelli, A. Ferri, Supercritical solvent impregnation of piroxicam on PVP at various polymer molecular weights, Journal of Supercritical Fluids 49 (2009) 271-278. 20. K. Gong, R. Viboonkiat, I.U. Rehman, G. Buckton, J.A. Darr, Formation and characterization of porous indomethacin-PVP coprecipitates prepared using solvent-free supercritical fluid processing, J. Pharm. Sci. 94 (2005) 2583-2590. 21. L. Manna, M. Banchero, D. Sola, A. Ferri, S. Ronchetti, S. Sicardi, Impregnation of PVP microparticles with ketoprofen in the presence of supercritical CO2, Journal of Supercritical Fluids 42 (2007) 378-384. 22. S. Ugaonkar, A.C. Nunes, T.E. Needham, Effect of n-scCO(2) on crystalline to amorphous conversion of carbamazepine, Int. J. Pharm. 333 (2007) 152-161. 23. K. Gong, I.U. Rehman, J.A. Darr, Characterization and drug release investigation of amorphous drug-hydroxypropyl methylcellulose composites made via supercritical carbon dioxide assisted impregnation, J. Pharm. Biomed. Anal. 48 (2008) 1112-1119. 24. S.G. Kazarian, G.G. Martirosyan, Spectroscopy of polymer/drug formulations processed with supercritical fluids: in situ ATR-IR and Raman study of impregnation of ibuprofen into PVP, Int. J. Pharm. 232 (2002) 81-90. 25. S. Ugaonkar, T.E. Needham, G.D. Bothun, Solubility and partitioning of carbamazepine in a two-phase supercritical carbon dioxide/polyvinylpyrrolidone system, Int. J. Pharm. 403 (2011) 96-100. 26. M. Maniruzzaman, M.M. Rana, J.S. Boateng, J.C. Mitchell, D. Douroumis, Dissolution enhancement of poorly water-soluble APIs processed by hot-melt extrusion using hydrophilic polymers, Drug Dev. Ind. Pharm. 39 (2013) 218-227. 27. G. Terife, P. Wang, N. Faridi, C.G. Gogos, Hot melt mixing and foaming of soluplus (R) and indomethacin, Polym. Eng. Sci. 52 (2012). 28. D.E. Alonzo, G.G.Z. Zhang, D. Zhou, Y. Gao, L.S. Taylor, Understanding the Behavior of Amorphous Pharmaceutical Systems during Dissolution, Pharm. Res. 27 (2010) 608-618.

ACS Paragon Plus Environment

Page 22 of 41

Page 23 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

29. S.Y. Lin, K.S. Chen, H.S. Teng, Effect of protective colloids on the induction of polymorphic changes in indomethacin agglomerates after solvent evaporation from o/w emulsions, J. Microencapsul. 16 (1999) 39-47. 30. 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. 31. M. Sihvonen, E. Järvenpää, V. Hietaniemi, R. Huopalahti, Advances in supercritical carbon dioxide technologies, Trends Food Sci. Technol. 10 (1999) 217-222. 32. Z. Guo, M. Lu, Y. Li, H. Pang, L. Lin, X. Liu, C. Wu, The utilization of drug-polymer interactions for improving the chemical stability of hot-melt extruded solid dispersions, J. Pharm. Pharmacol. 66 (2014) 285-296. 33. L. Taylor, G. Zografi, Spectroscopic characterization of interactions between PVP and indomethacin in amorphous molecular dispersions, Pharm. Res. 14 (1997) 1691-1698. 34. K. Mishima, Biodegradable particle formation for drug and gene delivery using supercritical fluid and dense gas, Adv. Drug Deliv. Rev. 60 (2008) 411-432. 35. P. Chattopadhyay, R. Huff, B.Y. Shekunov, Drug encapsulation using supercritical fluid extraction of emulsions, J. Pharm. Sci. 95 (2006) 667-679. 36. S. Shah, A. Pal, R. Gude, S. Devi, A novel approach to prepare etoposide-loaded poly(N-vinyl caprolactam-co-methylmethacrylate) copolymeric nanoparticles and their controlled release studies, J Appl Polym Sci 127 (2013) 4991-4999. 37. S.W. Kuo, S.C. Chan, F.C. Chang, Miscibility enhancement on the immiscible binary blend of poly(vinyl acetate) and poly(vinyl pyrrolidone) with bisphenol A, Polymer 43 (2002) 3653-3660. 38. L.G. Parada, L.C. Cesteros, E. Meaurio, I. Katime, Miscibility in blends of poly(vinyl acetate-co-vinyl alcohol) with poly(N,N-dimethylacrylamide), Polymer 39 (1998) 1019-1024. 39. C. Huang, F. Chang, Comparison of hydrogen bonding interaction between PMMA/PMAA blends and PMMA-co-PMAA copolymers, Polymer 44 (2003) 2965-2974. 40. T. Matsumoto, G. Zografi, Physical properties of solid molecular dispersions of indomethacin with poly(vinylpyrrolidone) and poly(vinylpyrrolidone-co-vinylacetate) in relation to indomethacin crystallization, Pharm. Res. 16 (1999). 41. J.E. Patterson, M.B. James, A.H. Forster, R.W. Lancaster, J.M. Butler, T. Rades, Preparation of glass solutions of three poorly water soluble drugs by spray drying, melt extrusion and ball milling, Int. J. Pharm. 336 (2007) 22-34. 42. N.H. Anderson, M. Bauer, N. Boussac, R. Khan-Malek, P. Munden, M. Sardaro, An evaluation of fit factors and dissolution efficiency for the comparison of in vitro dissolution profiles, J. Pharm. Biomed. Anal. 17 (1998) 811-822. 43. D.D. Sun, P.I. Lee, Evolution of Supersaturation of Amorphous Pharmaceuticals: The Effect of Rate of Supersaturation Generation, Molecular Pharmaceutics 10 (2013) 43304346.

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

44. D.E. Alonzo, Y. Gao, D. Zhou, H. Mo, G.G.Z. Zhang, L.S. Taylor, Dissolution and Precipitation Behavior of Amorphous Solid Dispersions, J. Pharm. Sci. 100 (2011) 33163331. 45. H.R. Guzman, M. Tawa, Z. Zhang, P. Ratanabanangkoon, P. Shaw, C.R. Gardner, H. Chen, J. Moreau, O. Almarsson, J.F. Remenar, Combined use of crystalline salt forms and precipitation inhibitors to improve oral absorption of celecoxib from solid oral formulations, J. Pharm. Sci. 96 (2007) 2686-2702. 46. H. Guzmán, M. Tawa, Z. Zhang, P. Ratanabanangkoon, P. Shaw, P. Mustonen, C.R. Gardner, H. Chen, J.P. Moreau, O. Almarsson, A “spring and parachute” approach to designing solid celecoxib formulations having enhanced oral absorption (2004). 47. R. Vandecruys, J. Peeters, G. Verreck, M.E. Brewster, Use of a screening method to determine excipients which optimize the extent and stability of supersaturated drug solutions and application of this system to solid formulation design, Int. J. Pharm. 342 (2007) 168-175. 48. P. Kanaujia, G. Lau, Wai Kiong Ng, E. Widjaja, A. Hanefeld, M. Fischbach, M. Maio, R.B.H. Tan, Nanoparticle Formation and Growth During In Vitro Dissolution of Ketoconazole Solid Dispersion, J. Pharm. Sci. 100 (2011) 2876-2885. 49. F. Tres, K. Treacher, J. Booth, L.P. Hughes, S.A.C. Wren, J.W. Aylott, J.C. Burley, Real time Raman imaging to understand dissolution performance of amorphous solid dispersions., Journal of controlled release : official journal of the Controlled Release Society 188 (2014) 53-60. 50. M.F. Pina, M. Zhao, J.F. Pinto, J.J. Sousa, D.Q.M. Craig, The Influence of Drug Physical State on the Dissolution Enhancement of Solid Dispersions Prepared Via Hot-Melt Extrusion: A Case Study Using Olanzapine, J. Pharm. Sci. 103 (2014) 1214-1223. 51. D.Q.M. Craig, The mechanisms of drug release from solid dispersions in water-soluble polymers, Int. J. Pharm. 231 (2002) 131-144. 52. W. Sinclair, M. Leane, G. Clarke, A. Dennis, M. Tobyn, P. Timmins, Physical Stability and Recrystallization Kinetics of Amorphous Ibipinabant Drug Product by Fourier Transform Raman Spectroscopy, J. Pharm. Sci. 100 (2011) 4687-4699.

ACS Paragon Plus Environment

Page 24 of 41

Page 25 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Table 1. Extrusion conditions for formulations. Drug loading [%] 10 30 50 10 30 50

Polymer PVP

SOL

Extrusion temperature [°C] 160 150 140 140 135 125

Table 2. Thermal characteristics of the pure materials prior to- and post-scCO2 treatment measured using a heating rate of 200 ⁰C/min. Standard deviations with n=3 are shown in brackets. aTgs estimated using the half-width method.

Pre-processing

Post processing

INM ∆H [J/g]

112.68 (3.32)

113.09 (1.02)

INM MP [⁰C]

175.38 (0.33)

178.52 (2.15)

PVP Tg [⁰C]a

128.56 (1.45)

131.10 (1.78)

SOL Tg [⁰C]a

83.49 (1.34)

82.28 (1.27)

Table 3. SFI processing conditions examined for the production of INM-SOL ASDs indicating the method development to produce completely amorphous solid dispersions containing Soluplus.

Identifier Drug loading [% w/w] SOLa 20 SOLb 20 SOLc 20 SOLd 20

Temperature [⁰⁰C] 60 60 70 70

ACS Paragon Plus Environment

Pressure [MPa] 10.0 15.0 10.0 15.0

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 41

Table 4. Correlation factors were calculated using SpectrumIMAGE for the wavenumber range 1740-1545 cm-1, between each spectrum in the Raman map for each formulation and the spectrum for amorphous INM. Lowest and highest correlation factor as well as the range. The range of correlation factors is greater for SFI formulations than for the respective HME counterparts. Raman spectra correlation factors Formulation Highest and lowest

Range

10 PVP HME

0.980 – 0.966

0.014

10 PVP SFI

0.979 – 0.955

0.024

30 PVP HME

0.995 – 0.982

0.013

30 PVP SFI

0.985 – 0.968

0.017

10 SOL HME

0.981 – 0.969

0.012

10 SOL SFI

0.980 – 0.954

0.026

30 SOL HME

0.994 – 0.992

0.002

30 SOL SFI

0.988 – 0.975

0.013

Table 5. Physical and thermal properties used in the calculation of the glass transition temperature of amorphous solid dispersions predicted by the Gordon Taylor model. Density (g/cm3)

Tg (°C)

aINM

1.32

42.81

PVP

1.20

120.83

SOL

0.99

73.25

ACS Paragon Plus Environment

Page 27 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Table 6. Glass transition of formulations measured by MDSC. Standard deviations with n=3 are shown in brackets. aGlass transition value predicted by Gordon-Taylor bNot completely amorphous. Drug [%] 10 30 50 10 30 50

Polymer PVP

SOL

HME 100.58 (2.86) 79.17 (2.44) 67.54 (0.30) 64.67 (0.47) 55.56 (1.92) 51.07 (1.46)

Tg [⁰C] SFI G-Ta 102.45 (0.77) 103.52 82.16 (2.00) 79.96 67.85 (1.67)b 64.69 63.63 (0.67) 65.22 58.57 (0.86) 56.52 57.57 (0.30)b 50.88

Table 7. Correlation factor R2 for a linear relationship between drug concentration and time for each formulation.

Formulation

R2

10 PVP SFI

0.711

30 PVP SFI

0.907

50 PVP SFI

0.804

10 PVP HME

0.953

30 PVP HME

0.962

50 PVP HME

0.988

10 SOL SFI

0.975

30 SOL SFI

0.976

50 SOL SFI

0.893

10 SOL HME

0.899

30 SOL HME

0.971

50 SOL HME

0.897

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table 8. Enthalpy of melting measured using HyperDSC, heating rate 200 °C/min, following 4 months accelerated stability, storage conditions 40 ⁰C and 75 %RH. Standard deviations with n=3 are shown in brackets.

Formulation

∆H [J/g]

10 PVP SFI

6.87 (1.36)

10 PVP HME

2.61 (0.06)

30 PVP SFI

9.22 (2.61)

30 PVP HME

5.88 (1.38)

50 PVP HME

9.29 (2.98)

10 SOL SFI

0.00 (0.00)

10 SOL HME

0.09 (0.11)

30 SOL SFI

0.90 (0.65)

30 SOL HME

1.77 (0.81)

50 SOL HME

2.30 (0.60)

ACS Paragon Plus Environment

Page 28 of 41

Page 29 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 1. Schematic representation of supercritical fluid impregnation apparatus including pressure vessel and ancillary equipment. 256x134mm (300 x 300 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. X-ray diffractograms of INM a) before and b) after treatment with scCO2. 144x111mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 30 of 41

Page 31 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 3. FTIR spectra of (top) INM, (middle) PVP and (bottom) SOL, a) before and b) after treatment with scCO2. The increased intensity of the band at 2338 cm-1 indicates the presence of residual carbon dioxide which remained within the polymer following depressurisation. 150x213mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4. Powder X-ray diffractograms of Soluplus/drug combinations processed using different conditions. Process conditions are detailed in Table 3. 142x111mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 32 of 41

Page 33 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 5. PXRD diffractograms of a) 10 HME, b) 10 SFI, c) 30 HME, d) 30 SFI, e) 50 HME and f) 50 SFI indicating that an amorphous drug loading of 50% INM could be attained using HME for both PVP (top) and SOL (bottom) while only 30% was achieved using the SFI method. 120x162mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6. Raman spectra collected from a 0.16 mm2 area of (left) PVP and (right) SOL formulations illustrating a) γ-INM, b) aINM, c) 10 HME, d) 10SFI, e) 30 HME, f) 30 HME, g) 50 HME, h) 50SFI, i) pure polymer. 287x202mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 34 of 41

Page 35 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 7. Raman maps for collected for (top) 50 SOL HME and (bottom) 50 SOL SFI and the reference spectrum (aINM) using SpectrumIMAGE software. A sample of spectra collect from each spectrum is also shown. Higher correlation factors near 1 indicate a good similarity to aINM. 239x232mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 8. FTIR spectra illustrating the carbonyl stretching region of different drug loaded and processed (SFI versus HME) formulations 286x106mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 36 of 41

Page 37 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 9. DSC thermograms of (left) INM-PVP and (right) INM-SOL formulations a) 10 HME b) 10 SFI c) 30 HME d) 30 SFI e) 50 HME f) 50 SFI. A single Tg is observed for all formulations, indicated by a tick 284x120mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 10. Dissolution profile of a) PVP HME b) PVP SFI c) SOL HME and d) SOL SFI formulations in pH 1.2. 50% (λ), 30% (ν),10% (π), crystalline INM (ϒ). 225x205mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 38 of 41

Page 39 of 41

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 11. Graphical representation of AUC45min and AUC8hr for PVP (top) and SOL (bottom) formulations. * represents a statistical difference p