Effect of Formulation Additives on Drug Transport through Size

Jun 29, 2018 - The applied voltage was 40 kV, while the current was 30 mA. ... It consists of a donor and an acceptor chamber (20 mL volumes) ... From...
2 downloads 0 Views 4MB Size
Subscriber access provided by - Access paid by the | UCSB Libraries

Article

Effect of formulation additives on drug transport through size-exclusion membranes Enik# Borbás, Petra T#zsér, Konstantin Tsinman, Oksana Tsinman, Krisztina Takacs-Novak, Gergely Völgyi, Balint Sinko, and Zsombor Kristof Nagy Mol. Pharmaceutics, Just Accepted Manuscript • DOI: 10.1021/acs.molpharmaceut.8b00343 • Publication Date (Web): 29 Jun 2018 Downloaded from http://pubs.acs.org on June 30, 2018

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 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 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.

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 46 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

Effect of formulation additives on drug transport through size-exclusion membranes Enikő Borbás1, Petra Tőzsér1, Konstantin Tsinman2, Oksana Tsinman2, Krisztina TakácsNovák3, Gergely Völgyi3, Bálint Sinkó2, Zsombor K. Nagy1* 1

Department of Organic Chemistry and Technology, Budapest University of Technology and Economics, Budapest 1111, Hungary 2

3

Pion Inc., Billerica, Massachusetts 01821, United States

Department of Pharmaceutical Chemistry, Semmelweis University, Budapest 1092, Hungary

*Budapest 1111, Hungary, 3 Muegyetem rakpart, [email protected]

Abstract The aim of this research was to investigate the driving force of membrane transport through sizeexclusion membranes and to provide a concentration-based mathematical description of it to evaluate whether it can be an alternative for lipophilic membranes in formulation development of amorphous solid dispersions. Carvedilol, an anti-hypertensive drug, was chosen and formulated using solvent-based electrospinning to overcome the poor water solubility of the drug. Vinylpyrrolidone-vinyl acetate copolymer (PVPVA64) and Soluplus® were used to create two different amorphous solid dispersions of the API. The loaddependent effect of the additives on dissolution and permeation through regenerated cellulose membrane

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

Page 2 of 46

was observed by a side-by-side diffusion cell, µFLUXTM. The solubilizing effect of the polymers was studied by carrying out thermodynamic solubility assays. The supersaturation ratio (SSR, defined as the ratio of dissolved amount of the drug to its thermodynamic solubility measured in exactly the same medium) was found to be the driving force of membrane transport in the case of size-exclusion membranes. Although the transport through lipophilic and size-exclusion membranes are mechanistically different, in both cases the driving force of membrane transport in the presence of polymer additives was found to be the same. This finding may enable the use of size-exclusion membranes as an alternative to lipid membranes in formulation development of amorphous solid dispersions. KEYWORDS dissolution-permeation, size-exclusion membrane, transport, carvedilol, supersaturation ratio, driving force, flux, microflux ABBREVIATIONS CAR, Carvedilol; ASD, amorphous solid dispersion; API, active pharmaceutical ingredient; BCS, biopharmaceutical classification system; SSR, supersaturation ratio; DSC, differential scanning calorimetry; SEM, scanning electron microscopy; XRD, X-ray powder diffraction

1. INTRODUCTION The dissolution enhancement of BCS II classified (poorly water-soluble, but highly permeable) active pharmaceutical ingredients (APIs) have recently become the focus of attention on the field of formulation development.1,2 Although in vitro dissolution tests are most commonly used to characterize drug formulation in the development phase, the results of this test do not always correlate with in vivo bioavailability.3 The lack

ACS Paragon Plus Environment

Page 3 of 46 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

of correlation is in many cases due to the fact that by using solubilizing agents as formulation additives, not only the solubility and dissolution of the drug, but also the partition coefficient changes between the aqueous and lipophilic phase and in this way the speed of membrane transport (permeability) is altered. This phenomenon is the so-called solubility-permeability interplay, described in detail by Dahan & Miller4 and reported by various authors5–8. In many cases, it was demonstrated in vitro and in vivo as well that the use of cyclodextrins to increase the solubility of an API leads to the decrease in the apparent permeability.9–11 Moreover, the decrease in the apparent permeability was found to be proportional to the solubility enhancement achieved by the complexing agent.12,13 Although a similar interplay between solubility and permeability of an API was already discovered a long time ago in vitro in the case of surfactants14–17, the first in vivo data on the effect of surfactants was only published in 201518. Not only does the complexation or micellar solubilization of the API lead to decreased permeability, but so does the use of hydrotropic agents like co-solvents, or PEG-40019–21. While solubilizing additives yield an improvement in the aqueous solubility, this comes with a decrease in permeability; it seems that the amorphous form of the API has the ability to overcome this trade-off. The amorphous state of the API leads to an increased apparent solubility, higher than the equilibrium crystalline solubility,22,23 meaning supersaturated solutions are created when dissolving amorphous drugs. Increased fluxes, compared to results from saturated or subsaturated solutions, were reported in many cases when supersaturated solutions were used5,24–26. These results confirm that amorphous solid dispersions (ASDs) have a great potential to improve the bioavailability of poorly water-soluble drugs. Electrospinning (ES) is a flexibly scalable and continuous technology to generate ASDs27–34.No matter how many working fluids are treated simultaneously, the generated fibers-based ASDs are always collected in the physical forms of non-woven membranes.35,36 Polymer additives, like PVP derivatives8 commonly used for electrospinning or any other formulation method producing ASDs, generally do not

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

Page 4 of 46

alter the equilibrium crystalline solubility of APIs.37 Instead polymers are responsible for making the API’s supersaturation last longer in solution by inhibiting precipitation and crystal growth.38 In some special cases polymers can even modify the solubility, like the non-ionic surfactant Soluplus8,39. The flux results of ASDs and solubilizing-additive-containing formulations clearly show in vivo and in vitro (Caco-2 and lipid membranes) that the driving force of membrane transport cannot be simplified to the concentration gradient, especially in cases where solubility is altered. The driving force of membrane transport in the case of lipophilic membranes was found to be the SSR (defined as the ratio of dissolved amount of the drug to its thermodynamic solubility measured in exactly the same medium with the same excipient concentration).8 Surprisingly, in the case of studying flux across regenerated cellulose-based size-exclusion membranes, similar patterns in solubility and permeability were observed when using solubilizing agents on the donor side.40,41 This is an interesting phenomenon because the transport of pure API is mechanistically different through lipophilic and size-exclusion membranes. In the case of transport through a lipid membrane, there are two aqueous phases separated by an apolar organic phase; this way the drug molecule first has to partition from the first aqueous phase into the organic phase to start diffusing there and then has to partition again into the acceptor phase.42 In the case of size-exclusion membranes, there is no partitioning process because the membrane, the donor and the acceptor side are filled with the same buffer solution; therefore, it measures a transport speed through an aqueous media, and not a organic phase like lipophilic membranes. The other important difference between the two types of membrane is how they distinguish the drug molecules. Only neutral species are able to efficiently cross the lipophilic membranes according to the pH-partition hypotheses43, while all ionized and neutral species are able to cross cellulose membranes, unless their transport is limited by their hydrodynamic size.

ACS Paragon Plus Environment

Page 5 of 46 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

Not just the mechanism but the application of the two membrane types differ as well. The celluloseacetate-based size-exclusion membranes are widely used in pharmaceutical research for characterization of the drug-release kinetics from microdisperse systems44 and for particle size measurement of aggregates, micelles or cyclodextrin complexes using different molecular weight cut-off membranes45. Recently, cellulose membranes were also applied in the description of the liquid-liquid phase separation phenomenon in highly supersaturated solutions of lipophilic drugs.24 Moreover, these membranes are used to model absorption of digested food in the intestines since they exclude macromolecules (like starch) and only let small digestive products (like maltose) permeate. A diffusion cell with lipid membrane aims to be predictive of passive permeation of the drug molecules through biological membranes. Therefore, use of biorelevant conditions, like the pH gradient, and tissue-specific lipid compositions30,46–48, have shown a good correlation with in vivo human intestinal absorption data47,49,50 in the case of pure APIs. Although size-exclusion membranes are permeable to ionized molecules and therefore cannot provide a good approximation of in vivo passive transcellular permeation properties in the case of pure APIs, unlike lipid membranes do, they might be a simple and sufficient tool to differentiate formulations containing different excipients but the same API. For that reason, the aim of this paper is to investigate the driving force of membrane transport for size-exclusion membranes and provide a concentrationbased mathematical description of it to evaluate whether it can be an alternative for lipophilic membranes in the case of studying drug formulations of the same API. Carvedilol (CAR) an anti-hypertensive drug was chosen as a poorly water-soluble model drug. It was processed to create an ASD by using solvent-based electrospinning and vinylpyrrolidone-vinyl acetate copolymer (PVPVA64) or Soluplus® as polymer additives. The load-dependent effect of the polymers was studied by simultaneous dissolution-permeation experiments and thermodynamic solubility assays.

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

Page 6 of 46

2. EXPERIMENTAL SECTION 2.1. Materials Carvedilol (406.482 g/mol, structure shown in Figure 1), buffer components (NaH2PO4, NaOH, KCl, NaCl, HCl solution) were purchased from Sigma-Aldrich Co. LLC. (St. Louis, MO, USA). Soluplus® (structure shown in Figure 1, average molecular weight 90 000–140 000 g/mol) and PVP VA64 (average molecular weight 24 000–30 000 g/mol were obtained from BASF (Ludwigshafen, Germany).

Figure 1. Structure of Carvedilol and the polymer additives

2.2. Electrospinning The polymer and CAR calculated amounts were added into 8 mL of the solvent (ethanol) and stirred magnetically (VELP Scientifica, Usmate, Italy) at 600 rpm until the dissolution was complete. The electrical potential applied on the spinneret electrode (inner diameter 0.8 mm) was 25 kV (NT-65 high voltage DC supply, MA2000 device, Unitronik Ltd, Nagykanizsa, Hungary). A grounded aluminum plate covered with aluminum foil was used as collector (20 cm from the spinneret). Polymer solutions were dosed with 4–8 mL/h at room temperature (25 °C) with an SEP-10S Plus type syringe pump (Aitecs, Vilnius, Lithuania). The composition of polymer solutions is shown in Table 1.

ACS Paragon Plus Environment

Page 7 of 46 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. Composition of the polymer solutions Ethanol Formulation

named

Polymer

after

(mL/g dry

API (w/w %) containing polymer

(w/w %) matter )

ES_PVP VA64

12

88

3.5

ES_Soluplus

12

88

3.5

2.3. Scanning Electron Microscopy The morphology of the samples was investigated with a JEOL JSM-6380LA (JEOL Ltd., Tokyo, Japan) type scanning electron microscope. Each specimen was fixed by conductive double-sided carbon adhesive tape and coated with gold-palladium alloy before examination. The applied accelerating voltage and working distance were 15 kV and 16 mm, respectively.

2.4. Differential Scanning Calorimetry Differential scanning calorimetry (DSC) measurements were carried out using a Setaram DSC 92 apparatus (Caluire, France). (Sample weight: 10–15 mg, open aluminum pan, nitrogen flush.) The temperature program consisted of an isothermal period (1 min) at room temperature, with subsequent linear heating from 25 °C to 150 °C at the rate of 10 °C/min.

2.5. X-ray powder diffraction X-ray diffraction patterns of the samples were recorded by means of a PANalytical (Amelo, the Netherlands) X’pert ProMDP X-ray diffractometer using Cu-K" α radiation (1.542 Å) and a Ni filter.

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

Page 8 of 46

The applied voltage was 40 kV, while the current was 30 mA. The samples were analyzed between 5° and 42° 2θ.

2.6. Measurement of the thermodynamic solubility with saturation shake flask method For measuring the thermodynamic solubility crystalline CAR Form I (5 mg) were added to 10 mL of 33 mM pH 6.5 phosphate buffer already containing the calculated amount of formulation additive (n = 3). The resulting mixtures were stirred at 25 °C for 6 h (to the solution equilibrium) followed by 18 h of sedimentation as recommended in the optimized protocol of shake flask method.51,52 Without filtration the concentration of the API in the saturated solution was determined by the immersed UV-probes of the µDiss (Pion Inc., Billerica MA, USA) apparatus based on a former calibration. For the evaluation of the concentrations, 5 mm tips and second derivative spectra were used. 2.7. In vitro dissolution-permeation study In vitro dissolution-permeation studies were carried out using a µFLUX (Pion Inc., Billerica MA, USA) apparatus (Figure 2). It consists of a donor and an acceptor chamber (20 mL volumes) separated by an artificial membrane selected to be a size-exclusion membrane for this study (Ultracel® regenerated cellulose, MWCO 1 kDa, 1.54 cm2, Merck KGaA, Darmstadt, Germany). At first, the membranes were soaked in the pH 6.5 phosphate (33 mM) buffer overnight, then 18 mL of pH 6.5 phosphate buffer was added into the donor and acceptor chambers. The pure drug powder or the formulations were placed into the donor chamber (n = 2–3). Both chambers were stirred at 150 rpm at 25 °C. In both chambers the API concentration was measured in real time by immersed UV-probes connected to the Rainbow instrument (Pion Inc., Billerica MA, USA). 5 mm pathlength tips on the donor and 20 mm pathlength tips on the acceptor probes were used. For the evaluation of the concentrations, second derivative spectra were used. The flux across the membrane was calculated using the following equation:

ACS Paragon Plus Environment

Page 9 of 46 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

J(t)=

where the flux (J) of a drug through the membrane is defined as the amount (n) of drug crossing a unit area (A) perpendicular to its flow per unit time (t).

Figure 2. Schematic figure of the µFLUXTM apparatus.

3. RESULTS AND DISCUSSION 3.1. Characterization of CAR containing formulations 3.1.1. Morphology ES_PVP VA64

ES_Soluplus

Figure 3. Scanning electron microscopic images of carvedilol containing electrospun formulations (PVP VA64 based on the left, Soluplus based on the right)

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

Page 10 of 46

Scanning electron microscopic (SEM) images show that with PVP VA64 beaded nano- and microscaled fibers were formed, while with Soluplus beadless nanoscaled electrospun fibers were obtained (Figure 3). Changing the polymer concentration can vary the solution viscosity and higher viscosity favors formation of fibers without beads.53 The sufficient polymer concentration to enable fiber formation differs for different polymers. The two polymers were used in the same concentration in the electrospinning solution (2g polymer/8 mL ethanol): with Soluplus this polymer concentration was high enough to achieve sufficient entanglements of the linear polymer in solution, this way enabling beadlesfiber formation54, while for PVP VA 64 the same polymer concentration was not enough in solution, this way only beaded fibers were produced. 3.1.2. X-ray powder diffraction

Figure 4. X-ray diffractograms of pure CAR (form II) and its formulations (a) CAR (form II), (b) physical mixture of CAR_PVP VA64 (12:88), (c) ES_PVP VA64, (d) physical mixture of CAR_Soluplus (12:88), (e) ES_Soluplus X-ray powder diffraction (XRD) was used to characterize the morphological changes of CAR induced by electrospinning. Figure 4 shows the diffractograms of the ES samples compared to the crystalline API and the physical mixtures. In the case of the physical mixtures (b,d), the characteristic peaks of the

ACS Paragon Plus Environment

Page 11 of 46 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

crystalline drug were clearly observed. While in the case of the formulation (c,e), no diffraction peaks were seen but only a diffuse background scattering as a proof of the amorphous nature of all the samples. 3.1.3. Differential Scanning Calorimetry

Figure 5. DSC thermograms of pure CAR (form II) and its formulations (a) CAR (form II), (b) physical mixture of CAR_PVP VA64 (12:88), (c) ES_PVP VA64, (d) physical mixture of CAR_Soluplus (12:88), (e) ES_Soluplus Differential scanning calorimetry analyses were also performed in order to study the physical state of the components in the samples. DSC thermograms (Figure 5) showed that CAR had a melting endotherm with a maximum at 114 °C; this melting peak could also be observed in the physical mixture. The absence of a melting peak in the CAR-containing electrospun formulations lead to the conclusion that the drug is in an amorphous form according to DSC results. Based on the SEM, XRD and DSC results dissolution enhancement was expected because of the high surface area and the amorphous state 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

Page 12 of 46

3.2 Thermodynamic solubility

Figure 6. CAR (form I) solubility in pH 6.5 phosphate buffer as a function of polymer additive concentration For the thermodynamic solubility studies, CAR form I was used because this polymorphic form was found to be more stable55 than form II, which was used in commercially available formulations. Compared to literature data on the solubility of form II at pH 6.556, the measured solubility of form I is around half of the solubility of the other polymorph. From the thermodynamic solubility results shown in Figure 6 it can be concluded that the addition of PVP VA64 to the dissolution media in the concentration range 0–600 µg/mL did not cause a significant enhancement in the thermodynamic solubility of the API. Applying the PVP VA64 in higher concentrations resulted in a gradual increase in the solubility of CAR. The addition of Soluplus showed a similar tendency of the solubility curve, except the solubilization of the API started at lower additive concentration. At 1500 µg/mL additive concentration, the solubility reached with Soluplus was found to be almost twice as high as the solubility in the presence of PVP VA64. 3.3. In vitro dissolution-permeability study on µFLUX™ platform

ACS Paragon Plus Environment

Page 13 of 46 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

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

Page 14 of 46

Figure 7. CAR concentration in the donor (left) and the acceptor (right) compartments ( (a) 16 µg/mL (b) 55 µg/mL (c) 130 µg/mL API load, pH 6.5 buffer in the donor and acceptor compartment, sizeexclusion membrane (1kDa MWCO), 150 rpm, 25 °C For the two formulations and the pure API, 8–12 different loads were applied with the theoretical maximum CAR concentration varying from 16 to 250 µg/mL in the donor chamber (for detailed flux data see Supporting information). From these the results of 3 different loads (16 µg/mL, 55 µg/mL and 100 µg/mL of CAR) are presented in Figure 7. They were chosen for the interpretation of the thermodynamic solubility results (Figure 6). Figure 7.a shows the results for 16 µg/mL CAR load, which also defines the polymer concentration to be 117 µg/mL. At this polymer concentration no significant solubility enhancement was observed in the case of either additive (Figure 6). With approximately the same solubilities and donor concentrations (Figure 7.a) the concentration-time profile in the acceptor chamber also shows very similar flux results for the pure API and both formulations (Table 2). Figure 7.b shows the dissolution-permeation results for 55 µg/mL CAR loads (403 µg/mL polymer load). In this concentration region of the polymers, Soluplus already solubilizes the API (Figure 6), while in the case PVP VA 64 solubilization is not yet significant. Such API concentration is well above its solubility

ACS Paragon Plus Environment

Page 15 of 46 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

at pH 6.5, so this concentration was achieved by pre-dissolving CAR in pH 1.6 solution (Carvedilol is a weak base with pKa 7.8, its solubility is pH dependent, and in acidic media is higher than in neutral or higher pHs56) and changing the pH of the solution to pH 6.5 with concentrated phosphate solution containing sodium hydroxide57. It can be seen from the acceptor results and from the calculated fluxes (Supporting information) that while the flux from the pure API and from the PVP VA64 containing ASD was found to be very similar, the flux from the Soluplus based ASD was found to be almost halved. Figure 7.c shows the results for 130 µg/mL load, meaning 953 µg/mL polymer concentration. At this polymer concentration, both PVP VA64 and Soluplus were found to act as solubilizers (Figure 6). Soluplus enhanced the solubility of the API to a greater extent and resulted in a more prominent decrease in flux than PVP VA64. In conclusion, the dissolution-permeation and solubility results show quite clearly that enhanced solubility (regardless of the mechanism of solubilization: micelles or secondary bonds) leads to decrease in flux results.

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

Page 16 of 46

4. MATHEMATICAL DESCRIPTION OF MEMBRANE TRANSPORT FOR SIZEEXCLUSION MEMBRANES 4.1 Description for the transport of the pure API

Figure 8. Transport model diagram, depicting two aqueous chambers separated by a size-exclusion membrane. The drug molecules are introduced in the donor as pure API. The transport of the pure API through porous membrane is most commonly described by Fick’s first law of diffusion (Eq.1), very similar to bulk diffusion, except that the diffusivity is called effective diffusivity to account for the longer diffusive pathway in porous media. =

(Eq.1)

where J is the flux, c0m and chm are the concentrations of the solute within the membrane at the two water-membrane boundaries (Figure 8) and where Dm is the effective diffusivity of the solute within the membrane. Considering that the concentration at the donor side of the membrane ( the bulk of the donor media (

) equals the concentration in

) (Eq.2) and the concentration at the acceptor side of the membrane (

)

equals the concentration in the bulk of the acceptor media ( ) (Eq.3) and substituting these into Eq.1,

ACS Paragon Plus Environment

Page 17 of 46 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

flux can be expressed as Eq.4. Separating the constant and variable parameters in Eq.4, Eq.5 was created, where the constants were merged into one parameter called the effective permeability (Pe). (Eq.2) (Eq.3) J=Dm

=Dm

J=

(Eq.4) =Pe

(Eq.5)2

From Eq.5, it can be seen that in the simple case when only the pure API is present, the driving force of membrane transport through the porous membrane is the concentration gradient. At the very beginning of the dissolution-permeation experiments, there was no API in the acceptor chamber. This was the reason why an assumption could be made that the concentration of the API was zero in the acceptor chamber, meaning the flux only depends on the concentration in the donor compartment: J=Pe

(Eq.6.)

Based on Eq.6, the flux results of pure API and CAR containing ASDs are shown in Figure 9 as a function of concentration reached in the donor chamber. The slope of lines are equal to the effective permeability.

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

Page 18 of 46

Figure 9. Flux of pure CAR and CAR containing ASDs at 25 °C as a function of concentration in the donor chamber It can be seen from Figure 9 that in the case of the pure API, the donor concentration was directly proportional to the flux up to 140 µg/mL. In the case of Soluplus containing ASD, direct proportionality between flux and the CAR concentration was only true for low Soluplus concentrations. The slope of the curve rapidly changed around 30 µg/mL CAR concentration, which equals 215 µg/mL in polymer concentration. This rapid decrease in the slope suggests that though the CAR concentration in solution increases the amount of API that is able to cross the membrane, it is not increasing with the same rate. This phenomenon was most probably caused by Soluplus micelles, solubilizing the API, but forming bigger aggregates than the molecular weight cut-off of the membrane (1 kDa). The polymer concentration at which the slope decrease occurred was in good agreement with the solubility results (Figure 6) where Soluplus showed to significantly enhance the solubility of the API in higher than 300 µg/mL additive concentrations. Although the increased solubility due to the addition of the non-ionic surfactant might seem a trivial phenomenon, in the case of Soluplus the effect on solubility seems rather API-specific since decrease in solubility of the API has been also reported8.

ACS Paragon Plus Environment

Page 19 of 46 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

PVP VA64 shows a similar effect on the flux-concentration curve only at higher CAR concentration (80 µg/mL) and accordingly higher polymer concentration (595 µg/mL). This polymer concentration was found to be in agreement with the solubility results, where higher than 600 µg/mL additive concentration could significantly enhance the solubility of the API. This phenomenon is surprising in the case of the PVP VA64 because its ability to form micelle-like structures in aqueous solution in the presence of API has not yet been reported in the literature to the author’s knowledge. Although the interaction between CAR and PVP VA64 in aqueous solution has not been yet investigated, the orienting effect of PVP through secondary bonds with CAR during controlled crystallization is a known phenomenon.58 In conclusion, Eq.6 was valid for the pure API, but only valid for a limited concentration range in the presence of polymer additives. The effective permeability of the API was found to be changing at higher additive concentration, as a result of solubilization of the API (due to micellization or secondary bond) in the donor chamber. These results indicate that Eq. 6 was only valid for the pure API, but could not describe the driving force in the presence of solubilizing additives. 4.2 Description for the transport of the API from ASDs containing polymers with high molecular weight

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

Page 20 of 46

Figure 10. Transport model diagram, depicting two aqueous chambers separated by a membrane barrier. The drug molecules are introduced in the donor chamber in formulation containing polymers instantly dissolving in the donor chamber, but not passing through the membrane. To find the universal driving force of membrane transport for size-exclusion membranes, independent from the type of formulation or additive, mathematical description is needed, which takes into account the solubility-permeability trade-off as well. The driving force can be derived from Fick’s first law (Eq.1). It is important to note that although in the case of size-exclusion membranes the donor and acceptor chamber and also the membrane pores are originally filled with the same media, when polymer additives are dissolved in the donor chamber and cannot cross the membrane because of their size, a modified donor media is created. In this media, the thermodynamic solubility of the API can differ from the solubility in the original buffer. The size-exclusion membrane separating these two solutions creates a phase boundary on the donor side of the membrane (Figure 10). In this case, the partitioning of the API between the solvents has to be considered (Eq.7). =

(Eq.7)

where Kd is the partition coefficient between donor compartment and membrane for steady state, cD is the concentration in the donor chamber, cm is the concentration in the membrane, * is the property at saturation. From Eq.7

can be expressed as:

(Eq.8) Considering that the concentration at the acceptor side of the membrane ( the bulk of the acceptor media ( ):

ACS Paragon Plus Environment

) equals the concentration in

Page 21 of 46 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

(Eq.9) If Eq.8 and Eq.9 is substituted to Eq.1: =

) (Eq.10)

Since the pores of the membrane are filled with the same media as the acceptor chamber, substituted with

can be

, resulting Eq.11. )=

) (Eq.11)

From Eq.11, it can be concluded that the driving force of membrane transport is the difference between the relative concentrations in the donor and acceptor compartment. This difference is directly proportional to the flux, and the coefficient of proportionality is

. This coefficient of

proportionality is a constant for a specific API when a specific membrane is used. This constant (B) can be defined for each API separately, like permeability, except B’s (Eq.10) value does not depend on the solubility of the API in donor and acceptor compartments. =

(Eq.12)

Considering that solubility in the donor could be influenced by formulation additives, Eq.13 could be used for expressing flux when measuring formulations and not just the pure API. J=B (

(Eq.13)

Usually in the beginning of the experiment, the relative concentration in the acceptor is negligible compared to

. For that reason, the flux results of CAR containing ASDs are shown in Figure 11 as a

function of only cD/cD*.

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

Page 22 of 46

Figure 11. Flux as a function of supersaturation ratio defined as the ratio of the donor concentration and the thermodynamic solubility of the API in the same media Figure 11 shows that the flux was directly proportional to SSR in the case of both formulations and the pure API. Comparing the slopes of lines shown in Figure 11 with homogeneity of slopes test, it can be concluded that the slopes of lines in the case of Soluplus, PVP VA64 containing formulations and the pure API did not differ significantly (p=0.5303). This means that the coefficient of proportionality is truly constant for Soluplus and PVP VA64 containing formulations.

ACS Paragon Plus Environment

Page 23 of 46 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

5. CONCLUSION While polymer additives generally did not act as solubilizing agents, Soluplus and PVP VA64 were found to increase the thermodynamic solubility of poorly water-soluble Carvedilol: the presence of 1467 µg/mL polymer resulted in 68.3 µg/mL solubility in the case of PVP VA 64 and 131 µg/mL solubility in the case of Soluplus, whereas the CAR concentration of saturated solution without additives was only found to be 37.7 µg/mL at pH 6.5. Although Soluplus was found to be almost twice as powerful as PVP VA64 in solubilizing CAR, the amount of drug permeated through the size-exclusion membrane was found to be significantly smaller in the case of Soluplus-containing formulations. These findings highlighted that simultaneous dissolution-permeation studies could be a useful tool for the selection and optimization of polymer matrixes for ASD-based formulations. Supersaturation ratio (SSR, defined as the ratio of dissolved amount of the drug to its thermodynamic solubility measured in exactly the same media) was found to be the driving force of membrane transport as a result of concentration-based mathematical description in the case of size-exclusion membranes. It was shown that although the transport through lipophilic and size-exclusion membranes seemed to be mechanistically different, in both cases the driving force of membrane transport in presence of additives was found to be the same. This finding may enable the use of size-exclusion membranes as an alternative to lipid membranes in formulation development (in the case that the additive cannot cross the membrane). SUPPORTING INFORMATION Flux results and donor concentration measured during in vitro dissolution-permeability study on µFLUX™ platform

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

Page 24 of 46

ACKNOWLEDGEMENT We thank our colleagues Dr. Zoltán Hórvölgyi and Dr. János Bódiss from the Department of Physical Chemistry and Material Science (Budapest University of Technology and Economics) who provided insight and expertise that greatly assisted the research. This project was supported by the New Széchenyi Plan (project ID: TÁMOP-4.2.1/B-09/1/KMR-2010-0002), Supported by the ÚNKP-17-3-I and the ÚNKP-17-4-II New National Excellence Program of the Ministry of Human Capacities, OTKA grants KH-124541 and PD-121051, FIEK_16-1-2016-0007, BME-KKP and the János Bolyai Research Scholarship of the Hungarian Academy of Sciences.

ACS Paragon Plus Environment

Page 25 of 46 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.

Keserü GM, Makara GM. The influence of lead discovery strategies on the properties of drug candidates. Nat Rev Drug Discov. 2009;8(3):203-212. doi:10.1038/nrd2796.

2.

Lipinski CA. Drug-like properties and the causes of poor solubility and poor permeability. J Pharmacol Toxicol Methods. 2000;44(1):235-249. doi:10.1016/S1056-8719(00)00107-6.

3.

Buckley ST, Fischer SM, Fricker G, Brandl M. In vitro models to evaluate the permeability of poorly soluble drug entities: Challenges and perspectives. Eur J Pharm Sci. 2012;45(3):235-250. doi:10.1016/j.ejps.2011.12.007.

4.

Dahan A, Beig A, Lindley D, Miller JM. The solubility–permeability interplay and oral drug formulation design: Two heads are better than one. Adv Drug Deliv Rev. 2016;101:99-107. doi:10.1016/J.ADDR.2016.04.018.

5.

Frank KJ, Rosenblatt KM, Westedt U, et al. Amorphous solid dispersion enhances permeation of poorly soluble ABT-102: True supersaturation vs. apparent solubility enhancement. Int J Pharm. 2012;437:288-293. doi:10.1016/j.ijpharm.2012.08.014.

6.

Frank KJ, Westedt U, Rosenblatt KM, et al. Impact of FaSSIF on the solubility and dissolution/permeation rate of a poorly water-soluble compound. Eur J Pharm Sci. 2012;47(1):16-20. doi:10.1016/j.ejps.2012.04.015.

7.

Avdeef A, Kansy M, Bendels S, Tsinman K. Absorption-excipient-pH classification gradient maps:

Sparingly

soluble

drugs

and

the

pH

partition

hypothesis.

2007.

doi:10.1016/j.ejps.2007.09.009. 8.

Borbás E, Sinkó B, Tsinman O, et al. Investigation and Mathematical Description of the Real Driving Force of Passive Transport of Drug Molecules from Supersaturated Solutions. Mol

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

Page 26 of 46

Pharm. 2016;13(11):3816–3826. doi:10.1021/acs.molpharmaceut.6b00613. 9.

Loftsson T, Brewster ME. Pharmaceutical applications of cyclodextrins: effects on drug permeation through biological membranes. J Pharm Pharmacol. 2011;63(9):1119-1135. doi:10.1111/j.2042-7158.2011.01279.x.

10.

Rao VM, Stella VJ. When can cyclodextrins be considered for solubilization purposes? J Pharm Sci. 2003;92(5):927-932. doi:10.1002/JPS.10341.

11.

Beig A, Agbaria R, Dahan A. Oral Delivery of Lipophilic Drugs: The Tradeoff between Solubility Increase and Permeability Decrease When Using Cyclodextrin-Based Formulations. Tajmir-Riahi H-A, ed. PLoS One. 2013;8(7):e68237. doi:10.1371/journal.pone.0068237.

12.

Beig A, Agbaria R, Dahan A. The use of captisol (SBE7-β-CD) in oral solubility-enabling formulations: Comparison to HPβCD and the solubility-permeability interplay. Eur J Pharm Sci. 2015;77:73-78. doi:10.1016/j.ejps.2015.05.024.

13.

Fine-Shamir N, Beig A, Zur M, Lindley D, Miller JM, Dahan A. Toward Successful Cyclodextrin Based Solubility-Enabling Formulations for Oral Delivery of Lipophilic Drugs: Solubility– Permeability Trade-Off, Biorelevant Dissolution, and the Unstirred Water Layer. Mol Pharm. 2017;14(6):2138-2146. doi:10.1021/acs.molpharmaceut.7b00275.

14.

Katneni K, Charman SA, Porter CJH. Impact of Cremophor-EL and Polysorbate-80 on Digoxin Permeability across Rat Jejunum: Delineation of Thermodynamic and Transporter Related Events Using

the

Reciprocal

Permeability

Approach.

J

Pharm

Sci.

2007;96(2):280-293.

doi:10.1002/JPS.20779. 15.

Amidon GE, Higuchi WI, Ho NFH. Theoretical and Experimental Studies of Transport of Micelle-Solubilized Solutes. J Pharm Sci. 1982;71(1):77-84. doi:10.1002/JPS.2600710120.

ACS Paragon Plus Environment

Page 27 of 46 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

16.

Molecular Pharmaceutics

Frank KJ, Westedt U, Rosenblatt KM, et al. Impact of FaSSIF on the solubility and dissolution/permeation rate of a poorly water-soluble compound. Eur J Pharm Sci. 2012;47(1):16-20. doi:10.1016/j.ejps.2012.04.015.

17.

Dahan A, Miller JM. The Solubility–Permeability Interplay and Its Implications in Formulation Design and Development for Poorly Soluble Drugs. AAPS J. 2012;14(2):244-251. doi:10.1208/s12248-012-9337-6.

18.

Hens B, Brouwers J, Corsetti M, Augustijns P. Gastrointestinal behavior of nano- and microsized fenofibrate: In vivo evaluation in man and in vitro simulation by assessment of the permeation potential. Eur J Pharm Sci. 2015;77:40-47. doi:10.1016/J.EJPS.2015.05.023.

19.

Yalkowsky SH, Valvani SC, Amidon GL. Solubility of nonelectrolytes in polar solvents IV: nonpolar

drugs

in

mixed

solvents.

J

Pharm

Sci.

1976;65(10):1488-1494.

doi:10.1002/JPS.2600651018. 20.

Riad LE, Sawchuk RJ. Effect of Polyethylene Glycol 400 on the Intestinal Permeability of Carbamazepine in the Rabbit. Pharm Res. 1991;8(4):491-497. doi:10.1023/A:1015803312233.

21.

Beig A, Miller JM, Dahan A. Accounting for the solubility–permeability interplay in oral formulation development for poor water solubility drugs: The effect of PEG-400 on carbamazepine

absorption.

Eur

J

Pharm

Biopharm.

2012;81(2):386-391.

doi:10.1016/J.EJPB.2012.02.012. 22.

Hancock BC, Parks M. What is the true solubility advantage for amorphous pharmaceuticals? Pharm Res. 2000;17(4):397-404.

23.

Alonzo DE, Zhang GGZ, Zhou D, Gao Y, Taylor LS. Understanding the behavior of amorphous pharmaceutical systems during dissolution. Pharm Res. 2010;27(4):608-618. doi:10.1007/s11095-

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

Page 28 of 46

009-0021-1. 24.

Raina SA, Zhang GGZ, Alonzo DE, et al. Impact of Solubilizing Additives on Supersaturation and Membrane Transport of Drugs. Pharm Res. 2015;32:3350-3364. doi:10.1007/s11095-0151712-4.

25.

Beig A, Miller JM, Lindley D, et al. Head-To-Head Comparison of Different Solubility-Enabling Formulations of Etoposide and Their Consequent Solubility-Permeability Interplay. J Pharm Sci. 2015:2941-2947. doi:10.1002/jps.24496.

26.

Mesallati H, Umerska A, Paluch KJ, Tajber L. Amorphous Polymeric Drug Salts as Ionic Solid Dispersion

Forms

of

Ciprofloxacin.

Mol

Pharm.

2017;14(7):2209-2223.

doi:10.1021/acs.molpharmaceut.7b00039. 27.

Vigh T, Horváthová T, Balogh A, et al. Polymer-free and polyvinylpirrolidone-based electrospun solid dosage forms for drug dissolution enhancement. Eur J Pharm Sci. 2013;49(4):595-602. doi:10.1016/j.ejps.2013.04.034.

28.

Balogh A, Drávavölgyi G, Faragó K, et al. Plasticized Drug‐Loaded Melt Electrospun Polymer Mats: Characterization, Thermal Degradation, and Release Kinetics. J Pharm Sci. 2014;103(4):1278-1287. doi:10.1002/jps.23904.

29.

Vrbata P, Berka P, Stránská D, Doležal P, Lázníček M. Electrospinning of diosmin from aqueous solutions for improved dissolution and oral absorption. Int J Pharm. 2014;473(1):407-413. doi:10.1016/j.ijpharm.2014.07.017.

30.

Borbás E, Balogh A, Bocz K, et al. In vitro dissolution–permeation evaluation of an electrospun cyclodextrin-based formulation of aripiprazole using µFluxTM. Int J Pharm. 2015;491(1):180-189. doi:10.1016/j.ijpharm.2015.06.019.

ACS Paragon Plus Environment

Page 29 of 46 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

31.

Molecular Pharmaceutics

Yu D-G, Li J-J, Zhang M, Williams GR. High-quality Janus nanofibers prepared using three-fluid electrospinning. Chem Commun. 2017;53(33):4542-4545. doi:10.1039/C7CC01661A.

32.

Wang Q, Yu D-G, Zhang L-L, Liu X-K, Deng Y-C, Zhao M. Electrospun hypromellose-based hydrophilic composites for rapid dissolution of poorly water-soluble drug. Carbohydr Polym. 2017;174:617-625. doi:10.1016/J.CARBPOL.2017.06.075.

33.

Szabó E, Démuth B, Nagy B, et al. Scaled-up preparation of drug-loaded electrospun polymer fibres and investigation of their continuous processing to tablet form. Express Polym Lett. 2018;12(5):436-451. doi:10.3144/expresspolymlett.2018.37.

34.

Farkas B, Balogh A, Farkas A, et al. Medicated straws based on electrospun solid dispersions. Period Polytech Chem Eng. 2018;62(3):310-316. doi:10.3311/PPch.11931.

35.

Xu Y, Li J-J, Yu D-G, Williams GR, Yang J-H, Wang X. Influence of the drug distribution in electrospun gliadin fibers on drug-release behavior. Eur J Pharm Sci. 2017;106:422-430. doi:10.1016/J.EJPS.2017.06.017.

36.

Wang K, Liu X-K, Chen X-H, Yu D-G, Yang Y-Y, Liu P. Electrospun Hydrophilic Janus Nanocomposites for the Rapid Onset of Therapeutic Action of Helicid. ACS Appl Mater Interfaces. 2018;10(3):2859-2867. doi:10.1021/acsami.7b17663.

37.

Raghavan SL, Kiepfer B, Davis AF, Kazarian SG, Hadgraft J. Membrane transport of hydrocortisone acetate from supersaturated solutions; the role of polymers. Int J Pharm. 2001;221:95-105. doi:10.1016/S0378-5173(01)00673-1.

38.

Ilevbare G, Liu H, Edgar K, Taylor L. Inhibition of solution crystal growth of ritonavir by cellulose

polymers–factors

influencing

polymer

effectiveness.

CrystEngComm.

2012;14(20):6503-6514. http://pubs.rsc.org/en/content/articlehtml/2012/ce/c2ce25515d. Accessed

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

Page 30 of 46

March 3, 2016. 39.

Hughey JR, Keen JM, Miller DA, Kolter K, Langley N, McGinity JW. The use of inorganic salts to improve the dissolution characteristics of tablets containing Soluplus-based solid dispersions. Eur J Pharm Sci. 2013;48:758-766. doi:10.1016/j.ejps.2013.01.004.

40.

Raina SA, Zhang GGZ, Alonzo DE, et al. Enhancements and limits in drug membrane transport using supersaturated solutions of poorly water soluble drugs. J Pharm Sci. 2014. doi:10.1002/jps.23826.

41.

Indulkar AS, Mo H, Gao Y, Raina SA, Zhang GGZ, Taylor LS. Impact of Micellar Surfactant on Supersaturation and Insight into Solubilization Mechanisms in Supersaturated Solutions of Atazanavir. Pharm Res. 2017;34(6):1276-1295. doi:10.1007/s11095-017-2144-0.

42.

Avdeef A. Absorption and Drug Development: Solubility, Permeability, and Charge State. 2nd ed. United States: John Wiley and Sons; 2012.

43.

Avdeef A, Tsinman O. PAMPA--a drug absorption in vitro model 13. Chemical selectivity due to membrane hydrogen bonding: in combo comparisons of HDM-, DOPC-, and DS-PAMPA models. Eur J Pharm Sci. 2006;28(1-2):43-50. doi:10.1016/j.ejps.2005.12.008.

44.

Washington C. Drug release from microdisperse systems: a critical review. Int J Pharm. 1990;58(1):1-12. doi:10.1016/0378-5173(90)90280-H.

45.

Saokham P, Sá Couto A, Ryzhakov A, Loftsson T. The self-assemble of natural cyclodextrins in aqueous solutions: Application of miniature permeation studies for critical aggregation concentration

(cac)

determinations.

Int

J

Pharm.

2016;505(1-2):187-193.

doi:10.1016/J.IJPHARM.2016.03.049. 46.

Avdeef A, Strafford M, Block E, Balogh MP, Chambliss W, Khan I. Drug absorption in vitro

ACS Paragon Plus Environment

Page 31 of 46 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

model: filter-immobilized artificial membranes. 2. Studies of the permeability properties of lactones

in

Piper

methysticum

Forst.

Eur

J

Pharm

Sci.

2001;14(4):271-280.

http://www.ncbi.nlm.nih.gov/pubmed/11684401. Accessed December 4, 2016. 47.

Sinkó B, Garrigues TM, Balogh GT, et al. Skin–PAMPA: A new method for fast prediction of skin penetration. Eur J Pharm Sci. 2012;45(5):698-707. doi:10.1016/j.ejps.2012.01.011.

48.

Müller J, Esső K, Dargó G, Könczöl Á, Balogh GT. Tuning the predictive capacity of the PAMPA-BBB model. Eur J Pharm Sci. 2015;79:53-60. doi:10.1016/J.EJPS.2015.08.019.

49.

Sugano K, Hamada H, Machida M, Ushio H, Saitoh K, Terada K. Optimized conditions of biomimetic artificial membrane permeation assay. Int J Pharm. 2001. doi:10.1016/S03785173(01)00845-6.

50.

Tam KY, Avdeef A, Tsinman O, Sun N. The permeation of amphoteric drugs through artificial membranes - An in combo absorption model based on paracellular and transmembrane permeability. J Med Chem. 2010;53(1):392-401. doi:10.1021/jm901421c.

51.

Baka E, Comer JEA, Takács-Novák K. Study of equilibrium solubility measurement by saturation shake-flask method using hydrochlorothiazide as model compound. J Pharm Biomed Anal. 2008;46(2):335-341. doi:10.1016/J.JPBA.2007.10.030.

52.

Avdeef A, Fuguet E, Llinàs A, et al. Equilibrium solubility measurement of ionizable drugs – consensus recommendations for improving data quality. ADMET DMPK. 2016;4(2):117. doi:10.5599/admet.4.2.292.

53.

Fong H, Chun I, Reneker D. Beaded nanofibers formed during electrospinning. Polymer (Guildf). 1999;40(16):4585-4592. doi:10.1016/S0032-3861(99)00068-3.

54.

Shenoy SL, Bates WD, Frisch HL, Wnek GE. Role of chain entanglements on fiber formation

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

Page 32 of 46

during electrospinning of polymer solutions: good solvent, non-specific polymer–polymer interaction limit. Polymer (Guildf). 2005;46(10):3372-3384. doi:10.1016/j.polymer.2005.03.011. 55.

Pataki H, Csontos I, Nagy ZK, et al. Implementation of Raman Signal Feedback to Perform Controlled Crystallization of Carvedilol. Org Process Res Dev. 2013;17(3):493-499. doi:10.1021/op300062t.

56.

Hamed R, Awadallah A, Sunoqrot S, et al. pH-Dependent Solubility and Dissolution Behavior of Carvedilol—Case Example of a Weakly Basic BCS Class II Drug. 2015. doi:10.1208/s12249015-0365-2.

57.

Borbás E, Nagy ZK, Nagy B, et al. The effect of formulation additives on in vitro dissolutionabsorption profile and in vivo bioavailability of telmisartan from brand and generic formulations. Eur J Pharm Sci. 2018;114:310-317. doi:10.1016/j.ejps.2017.12.029.

58.

Pataki H, Soti P, Vigh T, et al. Controlled Formation of Free-Flowing Carvedilol Particles in the Presence

of

Polyvinylpyrrolidone.

Chem

Eng

doi:10.1002/ceat.201300203.

ACS Paragon Plus Environment

Technol.

2014;37(2):249-256.

Page 33 of 46 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. Structure of Carvedilol and the polymer additives 70x78mm (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. Schematic figure of the µFLUXTM apparatus 60x45mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 34 of 46

Page 35 of 46 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. Scanning electron microscopic images of carvedilol containing electrospun formulations (PVP VA64 based on the left, Soluplus based on the right) 59x22mm (600 x 600 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. X-ray diffractograms of pure CAR (form II) and its formulations (a) CAR (form II), (b) physical mixture of CAR_PVP VA64 (12:88), (c) ES_PVP VA64, (d) physical mixture of CAR_Soluplus (12:88), (e) ES_Soluplus 70x62mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 36 of 46

Page 37 of 46 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. DSC thermograms of pure CAR (form II) and its formulations (a) CAR (form II), (b) physical mixture of CAR_PVP VA64 (12:88), (c) ES_PVP VA64, (d) physical mixture of CAR_Soluplus (12:88), (e) ES_Soluplus 67x56mm (600 x 600 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. CAR (form I) solubility in pH 6.5 phosphate buffer as a function of polymer additive concentration 77x75mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 38 of 46

Page 39 of 46 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. CAR concentration in the donor (left) and the acceptor (right) compartments ( (a) 16 µg/mL (b) 55 µg/mL (c) 130 µg/mL API load, pH 6.5 buffer in the donor and acceptor compartment, size-exclusion membrane (1kDa MWCO), 150 rpm, 25 °C 79x39mm (600 x 600 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 7. CAR concentration in the donor (left) and the acceptor (right) compartments ( (a) 16 µg/mL (b) 55 µg/mL (c) 130 µg/mL API load, pH 6.5 buffer in the donor and acceptor compartment, size-exclusion membrane (1kDa MWCO), 150 rpm, 25 °C 80x40mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 40 of 46

Page 41 of 46 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. CAR concentration in the donor (left) and the acceptor (right) compartments ( (a) 16 µg/mL (b) 55 µg/mL (c) 130 µg/mL API load, pH 6.5 buffer in the donor and acceptor compartment, size-exclusion membrane (1kDa MWCO), 150 rpm, 25 °C 79x39mm (600 x 600 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. Transport model diagram, depicting two aqueous chambers separated by a size-exclusion membrane. The drug molecules are introduced in the donor as pure API. 254x190mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 42 of 46

Page 43 of 46 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. Flux of pure CAR and CAR containing ASDs at 25 °C as a function of concentration in the donor chamber 74x70mm (600 x 600 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. Transport model diagram, depicting two aqueous chambers separated by a membrane barrier. The drug molecules are introduced in the donor chamber in formulation containing polymers instantly dissolving in the donor chamber, but not passing through the membrane. 254x190mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 44 of 46

Page 45 of 46 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. Flux as a function of supersaturation ratio defined as the ratio of the donor concentration and the thermodynamic solubility of the API in the same media 80x80mm (600 x 600 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

Graphical abstract 88x35mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 46 of 46