You are on page 1of 49

Subscriber access provided by Kaohsiung Medical University

Article
Phase Behavior of Drug-Hydroxypropyl Methylcellulose
Amorphous Solid Dispersions Produced from Various
Solvent Systems: Mechanistic Understanding of the Role
of Polymer using Experimental and Theoretical Methods
Naila Mugheirbi, Laura I. Mosquera-Giraldo, Carlos H. Borca, Lyudmila V. Slipchenko, and Lynne S. Taylor
Mol. Pharmaceutics, Just Accepted Manuscript • DOI: 10.1021/acs.molpharmaceut.8b00324 • Publication Date (Web): 06 Jun 2018
Downloaded from http://pubs.acs.org on June 7, 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 48 Molecular Pharmaceutics

1
2
3
4
5
Phase Behavior of Drug-Hydroxypropyl
6
7
8
9
Methylcellulose Amorphous Solid Dispersions
10
11
12
13
Produced from Various Solvent Systems:
14
15
16
17 Mechanistic Understanding of the Role of Polymer
18
19
20
21 using Experimental and Theoretical Methods
22
23
24
25
26
27
28
29 Naila A. Mugheirbi1, Laura I. Mosquera-Giraldo1, Carlos H. Borca2,3, Lyudmila V. Slipchenko2
30
31 and Lynne S. Taylor1*
32
33 1
Department of Industrial and Physical Pharmacy, College of Pharmacy, 575 Stadium Mall
34
35
36
Drive, Purdue University, West Lafayette, Indiana, 47907, United States.
37 2
38 Department of Chemistry, College of Science, Purdue University, 560 Oval Drive, West
39
40 Lafayette, Indiana, 47907, USA
41
42 3
Current Address: School of Chemistry and Biochemistry, College of Sciences, Georgia Institute
43
44
45 of Technology, 901 Atlantic Drive, Atlanta, Georgia 30332, USA
46
47 *To whom correspondence should be addressed: lstaylor@purdue.edu
48
49
50
51
52 KEYWORDS: Itraconazole, Alkyl alcohols, Evaporation rate, Amorphous solid dispersion, Spin
53
54 coating.
55
56
57
58 1
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 2 of 48

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 2
59
60 ACS Paragon Plus Environment
Page 3 of 48 Molecular Pharmaceutics

1
2
3 Abstract
4
5
6 The vast majority of studies evaluating amorphous solid dispersions (ASDs) utilize solvent
7
8 evaporation techniques as the preparation method. However, the impact of the solvent/co-solvent
9
10 system properties on the polymer conformation and the phase behavior of the resultant drug-
11
12
13
polymer blends is poorly understood. Herein, we investigate the influence of solvent properties
14
15 on the phase behavior of ASDs containing itraconazole (ITZ) and hydroxypropylmethyl
16
17 cellulose (HPMC) prepared using spin coating from binary/ternary co-solvent systems containing
18
19
alkyl alcohols, dichloromethane (DCM) and water. The compatibility of the polymer with the co-
20
21
22 solvent system was probed using high resolution imaging techniques supported by molecular
23
24 dynamics simulations. Solvent evaporation and evaporation rate profiles were tracked
25
26 gravimetrically to understand the impact of solvent composition on the evaporation process.
27
28
29 Short chain alcohols, including methanol (MeOH) and ethanol (EtOH) were found to induce
30
31 drug-polymer de-mixing in the presence of water, with EtOH being less sensitive to moisture
32
33 than MeOH owing to its ability to form an azeotrope with water. In contrast, water-induced
34
35
36
mixing was observed when higher alcohols, including n-propanol (PrOH) and n-butanol
37
38 (BuOH), were used as a co-solvent, due to the improved solubility of HPMC in the higher
39
40 alcohols in the presence of water. Isopropanol (IPA) produced phase separated ASDs under wet
41
42
and dry conditions with an increase in miscibility with faster evaporation rates in the presence of
43
44
45 water. This solvent-triggered phase behavior highlights the importance of conducting a thorough
46
47 screening of various solvents prior to the preparation of ASDs via solvent evaporation
48
49 approaches such as spray drying.
50
51
52
53
54
55
56
57
58 3
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 4 of 48

1
2
3 1. Introduction
4
5
6 Formulations containing active pharmaceutical ingredients (APIs) in amorphous form provide
7
8 great potential to overcome bioavailability issues associated with poorly soluble APIs. The
9
10 solubility advantage of the amorphous form of an API is attributed to their higher free energy,
11
12
13
and subsequently improved dissolution relative to crystalline forms, which can result in the
14
15 formation of a supersaturated solution.1 However, amorphous drugs tend to convert to the
16
17 thermodynamically stable crystalline state with time.2 To inhibit crystallization from the solid
18
19
and solution states, amorphous drugs are usually blended with a suitable polymer, producing
20
21
22 what is termed an amorphous solid dispersion (ASD).3
23
24 The physical stability of a given amorphous solid dispersion is highly dependent on the solubility
25
26 of the crystalline form and/or the miscibility of the amorphous phase of the drug in the polymeric
27
28
29 matrix. That is to say, if the drug loading is lower than the crystalline solubility, the system is
30
31 thermodynamically stable, and therefore no crystallization will take place. However, such low
32
33 drug loadings are typically impractical.4, 5
On the other hand, if the drug loading exceeds the
34
35
36
amorphous miscibility limit, amorphous-amorphous phase separation (AAPS) is
37
38 thermodynamically favored, leading to the formation of a drug-rich phase and a polymer-rich
39
40 phase. As a result, crystallization can take place in the drug-rich domains, which will ultimately
41
42
compromise any solubility and dissolution rate enhancements of the formulation.6
43
44
45 Understanding the phase behavior of polymer-polymer and drug-polymer blends has attracted a
46
47 great deal of interest over the past decade. Most studies conducted to understand phase behavior
48
49 of blends have employed solvent evaporation techniques for fabrication. However, only a limited
50
51
52 number of these studies has focused on the impact of the compatibility of the solvent system with
53
54 the polymer on the phase behavior of the resultant blend; the majority of work in this area has
55
56
57
58 4
59
60 ACS Paragon Plus Environment
Page 5 of 48 Molecular Pharmaceutics

1
2
3 focused on polymer mixtures.7-9 The substantial difference in physicochemical properties
4
5
6 between drug and polymer and the subsequent solubility gap between the two components in a
7
8 given solvent system is expected to induce solvent-triggered variations in phase behavior.10 Such
9
10 solvent-induced phase variation is more complicated for ASDs prepared from multicomponent
11
12
13
solvent systems as compared to those prepared from a single solvent system.11 However, the use
14
15 of more than one solvent is common during pharmaceutical processing. In this respect,
16
17 understanding the correlation between solvent composition, polymer conformation, and blend
18
19
miscibility is of utmost importance for drug-polymer systems. While some studies have
20
21
22 evaluated the influence of solvent composition on drug-polymer miscibility using techniques
23
24 such as differential scanning calorimetry (DSC) and polarized light microscopy (PLOM),8 these
25
26 are low resolution techniques which may lead to overestimation of drug-polymer miscibility.12
27
28
29 The itraconazole (ITZ)-hydroxypropyl methyl cellulose (HPMC) system has been reported to be
30
31 phase separated when produced via solvent evaporation from DCM:MeOH under uncontrolled
32
33 humidity conditions.12 Recently, we have correlated the origin of the reported phase separation to
34
35
36
the presence of water during the ASD production step.13 The objective of this study was to carry
37
38 out a thorough investigation of the impact of initial solvent composition on the polymer
39
40 conformation, and as a result the phase behavior of ITZ-HPMC ASDs produced using spin
41
42
coating. To this end, five alkyl alcohols, methanol (MeOH), ethanol (EtOH), isopropanol (IPA),
43
44
45 n-propanol (PrOH) and n-butanol (BuOH), were used as co-solvents with DCM. High resolution
46
47 characterization techniques were employed to investigate the phase behavior of the resultant
48
49 amorphous films and to understand the underlying phenomena controlling the behavior of each
50
51
52 solvent system. 13-15
53
54
55
56
57
58 5
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 6 of 48

1
2
3 2. Materials and Experimental Section
4
5
6 2.1. Materials
7
8 Itraconazole (ITZ) was purchased from Attix Pharmaceuticals (Ontario, Canada), methanol,
9
10 ethanol, isopropanol, n-propanol, n-butanol and dichloromethane were purchased from Fischer
11
12
13
Chemicals (NJ, U.S.A.). METHOCEL E3 grade was a gift from Dow Chemical Co. (Midland,
14
15 MI).
16
17 2.2.Preparation of Drug-Polymer Stock Solutions.
18
19
A set of stock solutions containing ITZ-HPMC in different solvents was prepared by dissolving a
20
21
22 30:70 w/w ITZ:HPMC blend in a 50:50 v/v mixture of each alkyl alcohol [MeOH, EtOH, IPA,
23
24 n-PrOH and n-BuOH] and DCM to have 10 mg/mL total solid content in solution.
25
26 2.3. Preparation of Spin Coated films.
27
28
29 A 100 µL aliquot of solution was spin coated on various substrates via dynamic dispensing using
30
31 a two-step procedure (500 rpm for 3 seconds and 2000 rpm for 30 seconds as reported previously
32
33 13
) at 21, 30, 35, 40 and 50 °C using a heat assisted spin coating set up. A heat lamp, served as a
34
35
36
heat source, and was placed above the sample zone of a KW-4A spin coater (Chemat
37
38 Technology Inc., CA, USA) and the temperature was monitored using an infrared thermometer
39
40 gun (VWR, Randor, Pa) with environmental relative humidity (RH) maintained at 50 ± 5% RH.
41
42
A two-step, slower spinning speed was also assessed (500 rpm for 15 seconds and 500 rpm for
43
44
45 15 seconds) at various temperatures; 21, 30, 35, 40 and 50 °C. Following spin coating, all
46
47 samples were stored under vacuum for at least 24 hours to ensure complete removal of the
48
49 solvent.
50
51
52 To investigate the texture of HPMC films produced under dry and wet conditions, a 7 mg/mL
53
54 solution of HPMC E3 solution in a 1:1 mixture of DCM:alkyl alcohol was prepared and filtered
55
56
57
58 6
59
60 ACS Paragon Plus Environment
Page 7 of 48 Molecular Pharmaceutics

1
2
3 using 0.22 PTFE filter. The solution was then divided into two samples where an aliquot (100
4
5
6 µL) of one sample was spin coated under a nitrogen purge (dry evaporation) and an aliquot of the
7
8 second sample was spin coated at 50% RH (wet evaporation). Spin coated HPMC films were
9
10 then left to dry overnight and the texture of these films was investigated using AFM.
11
12
13
2.4. Film Characterization
14
15 2.4.1. Fluorescence Microscopy
16
17 Spin coated films on a quartz substrate were imaged using an Olympus BX-51 fluorescence
18
19
microscope (NY, USA) equipped with an excitation filter to provide an excitation wavelength of
20
21
22 330-380 nm and an emission filter to permit emitted wavelengths from 420 nm onwards to be
23
24 detected. Exposure time was set to 0.2 seconds and ×100 objective lens was used for all samples.
25
26 2.4.2. Transmission Electron (TE) Microscopy
27
28
29 Electron diffraction and TE micrographs were acquired using a FEI Tecnai G 20 electron
30
31 microscope equipped with a LaB6 source and operated at 200 keV (Hillsboro, OR, USA).
32
33 Samples were prepared by spin coating onto 300 mesh carbon coated Copper TEM grids with a
34
35
36
standard thickness (5-6 nm) (SPI supplies, West Chester, PA), followed by incubation under
37
38 vacuum overnight to remove residual solvents.
39
40 2.4.3. Topographical Imaging
41
42
A NanoIR2 AFM-IR instrument (Anasys Instruments, Inc., Santa Barbara, CA) was used to
43
44
45 assess the topography of the spin coated films. Contact mode NIR2 probes (Model: PR-EX-
46
47 NIR2, Anasys Instruments, Inc., Santa Barbara, CA) were used to collect AFM topographical
48
49 images. A scan rate of 0.4 Hz was used for contact mode with an x and y resolution of 256
50
51
52 points. Topographical images were collected using the Analysis Studio software (version
53
54 3.10.5539. Lorentz contact resonance (LCR) spectroscopy and imaging was performed to
55
56
57
58 7
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 8 of 48

1
2
3 evaluate the distribution of drug and polymer within the film based on the variation in the contact
4
5
6 stiffness between the probe and the sample as described before.13 A Thermalever probe (Model:
7
8 EXP-AN2-300, Anasys Instruments Inc., Santa Barbara, CA) was used together with a small
9
10 LCR drive magnet to conduct LCR sweeps and collect height and LCR images. LCR sweeps
11
12
13
were then conducted to collect nanomechanical spectra of the area of interest. The spectra were
14
15 collected between 1 kHz and 200 kHz at 100 kHz/s scan rate. A drive strength of 20% was used
16
17 for LCR sweeps. In addition to nanomechanical spectra collection, LCR images were obtained
18
19
by driving the AFM cantilever at the contact resonance mode of one of the components. The
20
21
22 LCR drive strength was set to 50% to collect LCR images. The x and y resolution values were
23
24 set at 256 and 256 pt, and the scan rate was 0.3 Hz. Localized nanothermal (nanoTA)
25
26 measurements were also carried out. Temperature calibration was performed using three
27
28
29 crystalline polymers with melting points covering the entire temperature range of interest. The
30
31 polymers used were polycaprolactone (Tmelting = 55 °C), polyethylene (Tmelting = 116 °C), and
32
33 polyethylene terephthalate (Tmelting = 235 °C). A ThermaLever™ probe (model EXP-AN2-200,
34
35
36
Anasys Instruments Inc., Santa Barbara, CA) was used in the contact mode. The probe was
37
38 placed on the area of interest and the AFM tip was heated linearly with time, and the bending of
39
40 the probe was recorded. The temperature at which a thermal event, defined as penetration of the
41
42
probe into the surface of the sample due to sample softening, occurs is taken as an indication of
43
44
45 the glass transition for disordered substances.16
46
47 2.4.4. Determination of Temperature Dependent Solubility of Itraconazole
48
49 The solubility of ITZ in the investigated alcohols as a function of temperature was determined
50
51
52 using heat/cool cycles with temperature control achieved using a Crystal 16 Parallel Crystallizer
53
54 (Avantium, Amsterdam, Netherlands). Slurries of different ITZ concentrations (1-5.5 mg/ml) in
55
56
57
58 8
59
60 ACS Paragon Plus Environment
Page 9 of 48 Molecular Pharmaceutics

1
2
3 various alcohols were prepared in 1-dram clear glass vials. The vials were then placed in the
4
5
6 Crystal 16, heated to a given temperature (5 degrees below the alcohol boiling point) at 0.5
7
8 °C/min with continuous stirring at 300 rpm and the clear point was recorded. The temperature
9
10 was then reduced to 10 °C at 0.5 °C/min with the stirring maintained for the entire duration of
11
12
13
the experiment and the cloud point was recorded. Once the cloud point was reached, the
14
15 suspension was filtered using a Whatman 2.5 µm filter (Vernon Hills, IL, USA) and the collected
16
17 material was analyzed using DSC and PXRD.
18
19
2.4.5. Modulated Differential Scanning Calorimetry (MDSC)
20
21
22 Approximately 3-5 mg of a powdered sample was analyzed using a DSC Q2000 differential
23
24 scanning calorimeter (TA Instruments, Wood Dale, IL, USA) in a Tzero aluminium pan with a
25
26 crimped lid. Samples were heated from 25 to 200 °C using a 5 °C/min heating rate, ±0.796
27
28
29 modulation amplitude and a 60 s modulation period. The data were analyzed using Universal
30
31 Analysis® (version 4.7A) software.
32
33 2.4.6. Thermal Gravimetric Analysis
34
35
36
The evaporation profile of the stock solutions used for spin coating and the corresponding pure
37
38 solvent mixtures were measured using a TGA Q50 thermo-gravimetric analyzer (TA
39
40 Instruments, Wood Dale, IL, USA). Approximately 50 µL of each solution was placed into a
41
42
tared platinum pan immediately prior to placement inside a temperature-controlled furnace (24 ±
43
44
45 0.5°C) constantly purged with nitrogen at 60 mL/min flow rate. The weight loss was recorded
46
47 every minute until complete evaporation of the solvent occurred, which was confirmed by the
48
49 absence of additional weight loss for at least 10 minutes. The derivative of percentage weight
50
51
52 loss as a function of time is presented as the evaporation rate of the system.
53
54 2.5. Dynamic Light Scattering
55
56
57
58 9
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 10 of 48

1
2
3 The hydrodynamic diameter of HPMC in the solvent mixture before and during evaporation
4
5
6 under dry conditions and 50% RH, was analyzed using a Nano-Zetasizer (Nano-ZS) from
7
8 Malvern Instruments (Westborough, MA) equipped with dispersion technology software as
9
10 described before.8 A backscatter detector was used, and the scattered light was detected at an
11
12
13
angle of 173°. A 7 mg/mL solution of HPMC E3 solution in a 1:1 mixture of DCM:alkyl alcohol
14
15 was filtered using 0.22 PTFE filter prior to DLS analysis. The solution was then divided into two
16
17 samples where one sample was left to evaporate under a nitrogen purge (dry evaporation) while
18
19
and the second sample was left to evaporate at 50% RH (wet evaporation). Both solutions, wet
20
21
22 and dry evaporation, were left to evaporate until approximately 50% of the original volume had
23
24 been evaporated. An aliquot (700 µL) of the solution was placed in a glass cuvette and the size of
25
26 polymer molecules/aggregates was measured. The reported values are the average of at least four
27
28
29 measurements. Peaks with intensities below 5% were excluded from the plot.
30
31 2.6. Powder X-ray Diffraction
32
33 Powder X-ray diffraction (PXRD) of the filtrate collected immediately following the observation
34
35
36
of the cloud point in the Crystal 16 experiments was carried out using a Rigaku SmartlabTM
37
38 diffractometer (Rigaku Americas, The Woodlands, TX) with a Cu-Kα radiation source (λ =
39
40 1.542 Å) and a D/tex ultra-detector. Glass sample holders were used, and diffraction patterns
41
42
were obtained from 5°- 40° 2θ at a scan speed of 10°/min and a step size of 0.02°. The voltage
43
44
45 and current used were 40 kV and 44 mA, respectively.
46
47 3. Theoretical Considerations
48
49 3.1.Hansen Solubility Parameters
50
51
52 The key consideration for selecting the proper solvent/solvent mixture for a polymer/polymer
53
54 mixture is solubility. The internal energy of each component, i.e. that of the solvent and the
55
56
57
58 10
59
60 ACS Paragon Plus Environment
Page 11 of 48 Molecular Pharmaceutics

1
2
3 polymer, is one of the factors impacting the solubility of the polymer in that solvent.
4
5
6 Theoretically, polymer solubility can be predicted using the three-dimensional Hansen solubility
7
8 parameters (HSP). The total solubility parameter, δ, of a polymer, as proposed by Hansen, can be
9
10 calculated as the sum of squares of the partial solubility parameters, as shown in equation (1)17
11
12
   
13  =  +  +  
. (1)
14
15
16
Where the partial solubility parameters consist of δdi accounting for dispersion or nonpolar
17
18 effects, δpi for polar effects, and δhi to express the hydrogen bonding nature of the species The
19
20 total solubility parameter of a polymer is a point located in three-dimensional space. This point is
21
22
the center of a sphere called Hansen solubility sphere with radius R0 referred to as the interaction
23
24
25 radius of that polymer (Figure 1).
26
27 Equation (1) is used to calculate solubility parameters of solvents. For solvent mixtures,
28
29 solubility parameters can be calculated using the following equation;
30
31
32   = Ф  + Ф  + ⋯ , (2)
33
34 where Ф is the volume fraction of a mixture component.
35
36 The distance between the total solubility parameter of the solvent and the total solubility
37
38
39 parameter of the polymer (Rps) can be calculated using the following equation 17;
40

41    
42  = 4 − 
 +  − 
 +  − 
  . (3)
43
44 Where the "p" terms correspond to the parameters of the polymer and the "s" terms to the
45
46
47 parameters of the solvent. If the Rps value is less than the radius of interaction (R0) for the
48
49 polymer, the solvent or solvent blend is expected to dissolve the polymer (green point in Figure
50
51 1) and vice versa (red point in Figure 1).18
52
53
54
55
56
57
58 11
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 12 of 48

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 Figure 1. A schematic illustration of the Hansen’s solubility sphere theory.
31 In many cases, improved solvent quality can be achieved by choosing a solvent closer to the
32
33 center of the volume of the solubility sphere for a given polymer resulting in improved
34
35
36 solubility.17, 19 This can be achieved by adding a sufficient amount of an additional solvent which
37
38 is located in the direction one wants to move the total solubility parameter of the solvent mixture
39
40 towards. That is, to move it towards the center of the polymer’s solubility sphere. This concept
41
42
43
was proposed by Hansen (1967) and is based on the assumption that no specific interactions take
44
45 place between the components, and that the solvent displacement is through additive solubility
46
47 parameters based on their volume fraction.17
48
49
Although, solubility parameters cannot be used to quantitatively predict the solubility of the
50
51
52 polymer in each solvent, they were used herein to assess the potential miscibility of the polymer
53
54 with each alcohol. The interaction radius and the HSP distance between the polymer and various
55
56
57
58 12
59
60 ACS Paragon Plus Environment
Page 13 of 48 Molecular Pharmaceutics

1
2
3 alcohols were calculated using the values provided on the Hansen’s solubility parameters
4
5
6 website.20 The HSP distance upon mixing various alcohols with various fractions of water was
7
8 also calculated based on volume fractions as described in equation (2) using the three-
9
10 dimensional solubility parameters reported before 21.
11
12
13
3.2.Molecular Dynamics Simulations
14
15
16 To gather information regarding the dynamics of a HPMC polymer chain in different solvent
17
18 compositions, atomistic molecular dynamics (MD) simulations were performed in the molecular
19
20 modeling package GROMACS 5.0, using the CHARMM classical force field.22-24
21
22
23 (1) A protein data bank (PDB) file containing structural information for the ethanol molecule
24
25
26 was obtained from Networks/Pajek. The PDB file for n-propanol was retrieved from
27
28 virtualchemistry.org.25 The PDB files of IPA and DCM were extracted from the NIST
29
30 Chemistry Webbook.
31
32
33 (2) The structures of ethanol, propanol, IPA, and DCM were submitted to the online
34
35 topology building tool SwissParam, to obtain their corresponding topology files in
36
37 GROMACS format for the CHARMM force field.26
38
39
(3) Simulation boxes with a dimension of 4.0 × 4.0 × 4.0 nm were filled with the solvent
40
41
42 molecules proportioned to achieve molar fraction compositions of (1) 100% alcohol, (2)
43
44 70:30 % alcohol:DCM, (3) 50:50 % alcohol:DCM, and (4) 55:30:15 %
45
46 alcohol:DCM:water. The alcohols studied included: ethanol, propanol, and IPA.
47
48
49 Simulations including only solvent and solvent with polymer were setup with the following
50
51
52 conditions: The Velocity-Verlet algorithm was used as the time integrator. Periodic boundary
53
54 conditions for the simulation boxes were employed with the Verlet cutoff scheme for neighbor
55
56
57
58 13
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 14 of 48

1
2
3 searching. Short-range electrostatic interactions were explicitly considered up to a cut-off of 1
4
5
6 nm, and the particle mesh Ewald (PME) method modeled long-range electrostatic interactions.27
7
8 Van der Waals interactions were calculated up to a cut-off of 1 nm.
9
10
11 (4) In preparation for the simulations, the solvent boxes underwent three equilibration stages:
12
13 (1) a microcanonical ensemble equilibration over 0.5 ns (NVE), (2) a canonical ensemble
14
15
16 equilibration over 1 ns (NVT), (3) and an isothermal-isobaric equilibration over 1 ns
17
18 (NPT) using a time step of 1 fs/step. The solvent compressibility values used to perform
19
20 the NPT simulations were: 10.0 × 10-5 bar-1 for IPA, 9.0 × 10-5 bar-1 for propanol, 11.2 ×
21
22
23
10-5 bar-1 for ethanol, and 9.9 × 10-5 bar-1 for DCM. The compressibility values for
24
25 mixtures of two solvents were calculated by multiplying the compressibility by the molar
26
27 fraction of its corresponding solvent.
28
29
30 The polymer simulations involved eight stages:
31
32
33 (1) The chemical structure resembling HPMC E3 LV (30% methoxyl (M), 10%
34
35
36
hydroxypropoxyl (P), 60% hydroxyl (H)) was sketched in HyperChem 8.0.28 The
37
38 polymer chain included 20 monomers (537 atoms) and the composition is described in
39
40 Table 1.
41
42
43 Table 1. Structural information for the polymer chain resembling the composition of HPMC E3
44 LV. The substituents are denoted as methoxyl (M), hydroxypropoxyl (P), and hydroxyl (H).
45
There are three sites that can be substituted in each of the 20 monomers. Sites are denoted as C2,
46
47 C3, and C6, according to the position in the HPMC monomer.
48 Monomer
49
50 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
51
52 Position
53
54 C2 H H M M H H H M H M H P H M H H H H M H
55
56
57
58 14
59
60 ACS Paragon Plus Environment
Page 15 of 48 Molecular Pharmaceutics

1
2
3 C3 H M H H M P H H H M M H H H P M H M H H
4
5 C6 H M P H H H M H P H H H M H M H M P M H
6
7
8
9
10 (2) An energy minimization of the polymer chain was performed, with the Polak-Ribière
11
12
13
(conjugate-gradient) algorithm, using a root-mean-square (RMS) gradient of 1.0 × 10−2
14
15 kcal Å−1 mol−1 as convergence conditions.
16
17 (3) The HPMC structure was submitted to the online topology building tool SwissParam26, to
18
19
obtain topology files in GROMACS format for the CHARMM force field.
20
21
22 (4) Next, the polymer chain was solvated using the previously equilibrated solvent boxes.
23
24 The solvated system included ∼15,000 solvent molecules in cubic boxes with edges of up
25
26 to 12.0 nm.
27
28
29 (5) The solvated-polymer system was first minimized using a steepest descent algorithm
30
31 (0.01 fs/step), and the minimization was stopped when the maximum force was less than
32
33 100.0 kJ mol−1 nm−1.
34
35
36 (6) Second, a NVT equilibration was performed for 5 ns (1 fs/step) using the v-rescale
37
38 thermostat, at a reference temperature of 294 K, with a coupling constant (tau-t) of 0.2 ps.
39
40 (7) Third, an NPT equilibration was carried out for 5 ns (1 fs/step) using the v-rescale
41
42
thermostat, at a reference temperature of 294 K, with a coupling constant (tau-t) of 1.0 ps;
43
44
45 and the Berendsen barostat at a reference pressure of 1 bar, with a coupling constant (tau-
46
47 p) of 1.0 ps.
48
49 (8) Finally, the production run was performed for 20 ns, at 1 fs/step, recording output to disk
50
51
52 every 10 ps, under the isothermal-isobaric (NPT) ensemble, at a reference temperature of
53
54
55
56
57
58 15
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 16 of 48

1
2
3 294 K, with a coupling constant (tau-t) of 1.0 ps; and with the Berendsen barostat at a
4
5
6 reference pressure of 1 bar, with a coupling constant (tau-p) of 1.0 ps.
7
8 (9) The trajectories from the first NPT equilibration and the production run were analyzed,
9
10 and variations in the radius of gyration (Rg), and radial distribution functions (RDF) in
11
12
13
the presence of various solvent compositions are reported. The RDF is computed between
14
15 all the atoms of the polymer and all the atoms of the solvent molecules.
16
17
18 Following the same equilibration and production stage steps, a different set of simulations were
19
20 performed, using four polymer chains, and alcohol-DCM or alcohol-DCM-water compositions.
21
22 The purpose of this set of simulations was to determine the effect of water on the dynamics of
23
24
25 the polymer chains.
26
27
28 4. Results
29
30 The phase behavior of ITZ-HPMC film produced via spin coating from a 50:50 MeOH:DCM
31
32 solvent mixture has been reported previously, with phase separation attributed to the presence of
33
34
35
water in the system.12, 13 Herein, we examine the effect of the alkyl alcohol chain length on the
36
37 polymer conformation and as a consequence on the phase behavior of the produced films.
38
39 4.1.Phase Behavior of Films Produced at 50% RH.
40
41
FLOM was utilized as an initial screening tool to assess miscibility of the drug-polymer systems
42
43
44 based on the distribution of ITZ fluorescence intensity; ITZ is autofluorescent. At room
45
46 temperature (21 °C ± 0.5) and 50% RH, films produced using MeOH, EtOH and IPA co-solvent
47
48 systems displayed clear phase separation, with regions of high fluorescence intensity indicating
49
50
51 the non-uniform distribution of the drug in the films, as shown in Figure 2. In contrast, the use of
52
53 n-PrOH and n-BuOH as co-solvents produced miscible drug-polymer films, as inferred from the
54
55 homogenous distribution of the fluorescence intensity. Upon temperature increase, the size of the
56
57
58 16
59
60 ACS Paragon Plus Environment
Page 17 of 48 Molecular Pharmaceutics

1
2
3 phase separated domains decrease in films spin coated from DCM:MeOH and DCM:EtOH.
4
5
6 Films prepared using DCM:IPA showed an absence of discrete domains when prepared above
7
8 30°C, however, the fluorescence intensity distribution was inhomogeneous, with the
9
10 microstructure being similar to that associated with spinodal decomposition. 12
For films spin
11
12
13
coated from DCM:PrOH and DCM:BuOH, the temperature at which films were produced did not
14
15 impact their phase behavior as assessed by FLOM.
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 Figure 2. FL optical micrographs of ITZ-HPMC spin coated films, prepared at different
50
temperatures and 50% RH, using different alkyl alcohols as a co-solvent (scale bar corresponds
51
52 to 20 µm).
53
54
55
56
57
58 17
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 18 of 48

1
2
3 Due to the limited spatial resolution of FLOM imaging, complementary techniques were utilized
4
5
6 to enable further investigation of changes in the size of phase separated domains at elevated
7
8 temperature where the domain sizes are below the diffraction limit of optical microscopy. Thus,
9
10 TEM was employed as a high-resolution technique to investigate film texture. Bright field (BF)
11
12
13
TE microscopy enables phase separated domains in the nanometer range to be detected.12
14
15 Samples exhibiting the clearest evidence of phase separation, based on FLOM imaging, are
16
17 shown in Figure 3. BF TE micrographs shown in Figures 3a, b, c and d illustrate the reduction in
18
19
size of the discrete ITZ phase separated domains with an increase in temperature for samples
20
21
22 spin coated from DCM:MeOH and DCM:EtOH. The presence of a second form of phase
23
24 separation in the continuous phase of films produced at 21 and 50 °C from DCM:MeOH system
25
26 is also apparent; this was not seen in the DCM:EtOH films where only discrete domains were
27
28
29 observed (Figures 3c and d). Using DCM:BuOH at 21 °C led to homogeneous films, (Figure 3e),
30
31 confirming miscibility, in spite of the high RH conditions employed (50% RH). Films produced
32
33 from the IPA-containing system at various temperatures are displayed in Figures 3f-j. Films
34
35
36
produced at 21 and 30 °C displayed evidence of both discrete domains and a finely dispersed
37
38 microstructure, with the discrete domains being smaller when spin coating was carried out at
39
40 30°C. Upon increasing the evaporation temperature to 35 °C, fewer discrete domains were
41
42
observed with a change in the microstructure of the background. Increasing the evaporation
43
44
45 temperature to 50 °C eliminated the discrete domains and films appeared more homogeneous
46
47 than those produced at lower temperatures.
48
49
50
51
52
53
54
55
56
57
58 18
59
60 ACS Paragon Plus Environment
Page 19 of 48 Molecular Pharmaceutics

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Figure 3. BF TE micrographs of spin coated films at 50% RH from DCM:MeOH (a) at 21 and
23
(b) at 50 °C, DCM:EtOH (c) at 21 and (d) at 50 °C, DCM:BuOH (e) at 21 °C, DCM:IPA at (f)
24
25 21, (g) 30, (h) 35, (i) 40 and (j) 50 °C.
26
27 To investigate whether the reduction in discrete domain size at elevated temperature is due to a
28
29 temperature or an evaporation rate effect, a second set of films was produced at a lower spinning
30
31
speed (Figure SI.1). Films produced at a lower spinning rate exhibited larger domain sizes for
32
33
34 comparable temperatures suggesting that the reduction in domain size observed at high
35
36 temperatures is mainly due to an increase in evaporation rate.
37
38 4.2.Phase Behavior of Films Produced at 20% RH.
39
40
41 Because the phase separation observed for the ITZ:HPMC system spin coated from DCM:MeOH
42
43 is attributed to water, absorbed from the atmosphere during spin coating,12 it was of interest to
44
45 explore phase behavior for films prepared from the different alcohol co-solvent systems at a
46
47
48 lower environmental RH (20%). Films produced from DCM:MeOH solution, at 21 °C and 20%
49
50 RH (Figure 4a), were found to be phase separated in good agreement with previous
51
52 observations.12 In contrast, when EtOH was used as the co-solvent, FLOM and BF TEM images
53
54
indicated that a miscible system is produced (Figure 4b). Films from IPA were heterogeneous
55
56
57
58 19
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 20 of 48

1
2
3 but lacked discrete domains (Figure 4c). Somewhat surprisingly, PrOH and BuOH cosolvent
4
5
6 systems at low RH led to phase separated films (Figures 4d and e). Further reduction of the
7
8 environmental humidity to 7% RH did not produce miscible films in the case of DCM:IPA,
9
10 DCM:PrOH and DCM:BuOH systems as shown in Figures 4f, g, and h, respectively. These
11
12
13
results were counterintuitive based on observations that miscible films were produced from
14
15 MeOH:DCM at 7% RH.12 It should be highlighted that EtOH, IPA, PrOH and BuOH are capable
16
17 of forming an azeotropic mixture with water while MeOH cannot. This indicates that 95.5:4.5 %
18
19 29
EtOH:H2O system will behave as a one component system with a boiling point of 78.2 °C.
20
21
22 Similarly, 87.9:12.1 % IPA:H2O system will evaporate at 80.4 °C as a one component system.
23
24 Thus, when 20% RH was employed the water content, acquired from the environment, might
25
26 have been reduced to or below the azeotropic composition compared to the higher amount of
27
28
29 moisture absorbed when spin coating was carried out at 50% RH. PrOH and BuOH can form
30
29
31 71.7 and 55.5 % azeotropes with water, respectively. Thus, they can evaporate as a single
32
33 component solvent even with a larger amount of water which could explain the miscibility of the
34
35
36
films produced from those two systems at 50% RH.
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 20
59
60 ACS Paragon Plus Environment
Page 21 of 48 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 Figure 4. FL optical micrographs and BF TE micrographs of ITZ-HPMC spin coated films from
37 (a) DCM:MeOH, (b) DCM:EtOH, (c) DCM:IPA, (d) DCM:PrOH and (e) DCM:BuOH at 21 °C
38 and 20% RH and films spin coated from (f) DCM:IPA, (g) DCM:PrOH and (h) DCM:BuOH at
39 21 °C and 7% RH.
40
41
42 4.3.Investigation of Compositional Variation in the Phase Separated Films
43
44 Purohit and Taylor have previously identified the discrete domains as an ITZ-rich phase for
45
46 system prepared from DCM:MeOH under uncontrolled humidity conditions.12 Herein, it was of
47
48
49 interest to investigate the composition of the bi-continuous microstructure observed in some of
50
51 the films AFM coupled with Lorentz contact resonance (LCR) imaging was employed for this
52
53 purpose, focusing on films spin coated from DCM:IPA system at 20% RH. The topographical
54
55
images in Figure 5a revealed that the film surface was uneven with protruding features in a semi-
56
57
58 21
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 22 of 48

1
2
3 continuous arrangement. In LCR amplitude imaging, images are collected based on the relative
4
5
6 stiffness of components in the spin coated films upon applying the appropriate drive frequency.
7
8 As a result, individual components can be selectively highlighted where bright areas will be
9
10 visible when the contact frequency between the probe and the scanned area is the same as the
11
12
13
applied LCR drive frequency. Based on the nano-mechanical spectra in Figure 5c, ITZ is stiffer
14
15 than HPMC and thus, when the LCR drive frequency is set to 130 kHz, ITZ rich regions
16
17 appeared bright in the LCR amplitude image, while HPMC-rich regions are correspondingly
18
19
dark (Figure 5b). This indicates that the high features in the film, appearing as an orange color in
20
21
22 the 3D plot in Figure 5d, correspond to HPMC. These results were confirmed using nano-thermal
23
24 analysis of both bright and dark regions as can be seen in Figure 5e where the softening behavior
25
26 was investigated. The softening temperature of the high features was ~115 °C while that of the
27
28
29 continuous phase was ~60 °C consistent with the glass transition values of HPMC and ITZ,
30
31 respectively.12
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 22
59
60 ACS Paragon Plus Environment
Page 23 of 48 Molecular Pharmaceutics

1
2
3 Figure 5. AFM topographic image produced using contact mode (a) LCR amplitude image at
4
130 kHz (b) along with the nano-mechanical spectra collected for ITZ and HPMC (c), 3D
5
6 representation (d) and localized (nano-) thermal analysis (e) obtained from ITZ-HPMC films
7 spin coated from IPA:DCM at 21 °C and 20% RH in Figure 4a where the red and black marks
8 are located.
9
10
11 4.4. Impact of Water on the Evaporation Profile of Various Solvent Systems.
12
13
14 To better understand the solvent evaporation process and how it is impacted by the initial solvent
15
16 composition, the evaporation of solvent systems containing DCM with select alcohols, dry and in
17
18 the presence of a known amount of water, was monitored gravimetrically as described
19
20
21 previously.8 The evaporation profile of a given co-solvent mixture was found to be highly
22
23 dependent on the water content as shown in Figure 6. The DCM:MeOH mixture containing 4%
24
25 water showed the same initial evaporation rate as the dry solvent blend, but subsequently, the
26
27
evaporation rate was dramatically reduced such that the total evaporation time was extended by a
28
29
30 factor of 4. In contrast, the presence of 4% water does not seem to impact the evaporation rate of
31
32 the DCM:EtOH solvent system while the addition of 10% water to the mixture resulted in a 7-
33
34 fold increase in the evaporation time. Biphasic evaporation profiles were observed with dry
35
36
37 DCM:MeOH, dry DCM:EtOH and DCM:EtOH with 4% water with the evaporation rate
38
39 decreasing with time. More complex evaporation profiles were observed upon the addition of 4%
40
41 water to the DCM:MeOH and 10% water to the DCM:EtOH systems. Here, there was an initial
42
43
44 rapid decrease in the evaporation rate similar to that observed for the dry solvent blends followed
45
46 by a decreased evaporation rate. A sudden increase in the evaporation rate, in both samples, was
47
48 observed immediately prior to complete evaporation. For the DCM:IPA solvent system, the
49
50
addition of 4% water delayed complete evaporation by 2.5 minutes. However, the shape of the
51
52
53 evaporation and evaporation rate profiles were not affected. For DCM:PrOH and DCM:BuOH
54
55
56
57
58 23
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 24 of 48

1
2
3 solvent mixtures, 4% water did not affect the evaporation time compared to their dry
4
5
6 counterparts and the profiles were similar for dry and wet solvents.
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 24
59
60 ACS Paragon Plus Environment
Page 25 of 48 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 25
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 26 of 48

1
2
3 Figure 6. Evaporation profile (weight %) vs time and Evaporation rate profile (Weight change
4
%/min) vs time of pure solvent mixtures.
5
6
7
8 4.5.Solubility-Temperature Phase Diagram of ITZ in Alkyl Alcohols
9
10 The temperature dependent solubility line of ITZ in alkyl alcohols was determined using clear
11
12
13
and cloud point measurements for suspensions with different initial compositions. Figure SI.2
14
15 shows the “clear point” solubility line. Clear point is defined as the temperature at which all the
16
17 crystals are dissolved upon heating an ITZ crystalline suspension. Figure 7b shows the “cloud
18
19
point” solubility line. Cloud point is defined as the temperature at which turbidity appears once
20
21
22 the previously heated solution is cooled down.
23
24 Since the starting material is crystalline ITZ, the clear point corresponds to the crystalline
25
26 solubility of the drug. Crystalline ITZ shows the lowest solubility in IPA and the highest
27
28
29 solubility in MeOH. The properties of the particulate species following cooling the solution
30
31 were investigated using PXRD and DSC (Figures SI.3 and SI.4). The PXRD results indicated the
32
33 particulates formed during cooling were predominantly amorphous with some Bragg peaks
34
35
36
characteristic of crystalline ITZ being present for samples from MeOH, EtOH, IPA and PrOH;
37
38 these were absent in the BuOH precipitate. In agreement with these data, DSC thermograms
39
40 (reversing heat flow) for precipitates from lower alcohols exhibited a very weak glass transition
41
42
in addition to a melting endotherm characteristic for crystalline ITZ.30 In contrast, the precipitate
43
44
45 from the higher alcohols exhibited thermal events characteristic of glassy ITZ; a glass transition
46
47 (Tg) followed by two liquid crystalline transitions (TLC1 and TLC2).31 The phase separation
48
49 concentration during cooling is quite similar for all the alcohols tested except for IPA, where the
50
51
52 solubility in this alcohol is much lower than for the other systems evaluated.
53
54 4.6.An Investigation into the Polymeric Monomer/agglomerates Size
55
56
57
58 26
59
60 ACS Paragon Plus Environment
Page 27 of 48 Molecular Pharmaceutics

1
2
3 The sizes of HPMC species in stock solutions of different DCM:alcohols blends, before and after
4
5
6 partial evaporation, are shown in Figure 7. Acknowledging that the actual size is challenging to
7
8 measure and that only a comparison of the hydrodynamic diameters is carried out herein, values
9
10 below 10 nm are assigned to polymer molecules (individual chains) and are shown in Figure 7a,
11
12
13
while species above 10 nm are considered to indicate aggregates, and results are summarized in
14
15 Figure 7b. The mean hydrodynamic diameter of HPMC molecules measured in the initial dry
16
17 stock solutions before evaporation was 4 - 5 nm. There were no other peaks observed in the
18
19
MeOH and EtOH solutions with DCM. However, a second population with a mean
20
21
22 hydrodynamic diameter higher than 40 nm was observed in DCM solutions with the other
23
24 alcohols indicating that HPMC aggregates are also present in these systems. Following partial
25
26 evaporation under dry conditions, HPMC started to aggregate in all solutions albeit to different
27
28
29 degrees. Furthermore, the population of sub-10nm hydrodynamic radius species disappeared
30
31 from IPA, PrOH and BuOH solutions (green, brown and grey arrows respectively in Figure 7a).
32
33 Both molecules below 10 nm (Figure 7a) and some aggregates (Figure 7b) were observed in the
34
35
36
MeOH and EtOH solutions evaporated under dry conditions. Notably, the percentage of
37
38 aggregated HPMC in MeOH and EtOH solution upon evaporation under dry conditions was very
39
40 small (17 ± 4 and 11 ± 5%, respectively) with most of the polymer present as discrete chains.
41
42
In contrast, when solutions were evaporated at 50% RH, IPA was the only alcohol where
43
44
45 evaporation induced substantial aggregation of HPMC. A few small aggregates (~20 nm,
46
47 representing 6 ± 4% of the entire population) were observed in the solution of HPMC in EtOH
48
49 evaporated at 50% RH. Interestingly, in PrOH and BuOH mixtures with DCM, no aggregation
50
51
52 occurred upon evaporation at 50% RH. Rather, the aggregates observed in the original solutions
53
54 of DCM:PrOH and DCM:BuOH disappeared (brown and grey arrows, respectively).
55
56
57
58 27
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 28 of 48

1
2
3
4 Non hatched bars: Starting solution
5 Densly hatched bars: Dry evaporation
12 Sparsley Hatched bars: Wet evaporation
6

Percentage of the entire population (%)


7 100
8
10
Hydrodynamic diameter (nm)

9
10 80
11 8
12
13
60
14 6
15
16
17 4
40
18
19
20 2 20
21
22
23 0 0
24 MeOH EtOH IPA PrOH BuOH
25
26
27
28
29
30 Non hatched bars: Starting solution
31 Densly hatched bars: Dry evaporation
180 Sparsley Hatched bars: Wet evaporation 110
32

Percentage of the entire population (%)


33 100
160
34
Hydrodynamic diameter (nm)

35 90
140
36 80
37
120
38 70
39
100 60
40
41 50
80
42
43 40
60
44
30
45
40
46 20
47
20
48 10
49
0 0
50
51 MeOH EtOH IPA PrOH BuOH
52
53
54
55
56
57
58 28
59
60 ACS Paragon Plus Environment
Page 29 of 48 Molecular Pharmaceutics

1
2
3 Figure 7. Mean hydrodynamic diameter of (a) the molecules, and (b) aggregates of HPMC
4
alcohol-DCM initial solutions and solutions partially evaporated either under dry conditions or at
5
6 50% RH. The open triangles and second y-axis show the % intensity for the various species.
7
8
9 AFM investigation of the topography of the spin coated HPMC films from various co-solvent
10
11 systems under dry and wet conditions (Figure 8) could be rationalized by considering the data
12
13
14 obtained from the DLS study. Flat films with a smooth surface were produced when the polymer
15
16 presents as individual chains (molecules) including films spin coated from MeOH and EtOH
17
18 under dry and wet conditions and from PrOH and BuOH under wet conditions. In contrast, rough
19
20
21 films with irregular, corrugated surface topography, were produced when the solution contained
22
23 polymer aggregates including films produced from IPA under dry and wet conditions and from
24
25 PrOH and BuOH under dry conditions.
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 29
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 30 of 48

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 Figure 8. Schematic illustration of the correlation between the state of HPMC (individual
46 molecule vs aggregate) in the initial stock solution and upon evaporation under dry (20% RH)
47
48
and wet (50% RH) conditions, according to DLS data, and the texture of the HPMC spin coated
49 films investigated using AFM.
50
51 4.7.Solubility Parameters Calculations
52
53 The solubility parameters of Methocel have been reported previously by Archer (1992).18 The
54
55
56 reported values for the E3 grade employed in this study were, δdi= 17.4 MPa1/2, δpi= 14.9 MPa1/2
57
58 30
59
60 ACS Paragon Plus Environment
Page 31 of 48 Molecular Pharmaceutics

1
2
3 and δhi=19.3 MPa1/2. In addition, the radius of interaction (R0) was found to be 7.6 MPa1/2. The
4
5
6 distance between the total solubility parameters of the polymer and various alcohols employed in
7
8 this study is displayed in Figure 9a. The distances between the three-dimensional solubility
9
10 parameters center of HPMC and the total solubility parameters of MeOH and EtOH were less
11
12
13
than the interaction radius of HPMC (R0 represented by the dashed line in Figure 9a). In contrast,
14
15 the distances between the three-dimensional solubility parameters center of HPMC and IPA,
16
17 PrOH and BuOH were higher than the reported interaction radius (R0) of HPMC. These
18
19
observations point towards the decrease in HPMC solubility with an increase in alcohol chain
20
21
22 length. Nonetheless, it should be highlighted that the term solubility parameter does not
23
24 adequately represent the predictive solubility of the system. In addition, the assumption that no
25
26 non-additive interactions take place between various components of the system may also lead to
27
28 32
29 inaccuracies. However, this approach is useful to provide insight into the miscibility of a
30
31 polymer with a solvent, clearly indicating a decrease in polymer solubility with increasing
32
33 alcohol alkyl chain length.
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 31
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 32 of 48

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 32
59
60 ACS Paragon Plus Environment
Page 33 of 48 Molecular Pharmaceutics

1
2
3 Figure 9. (a) Hansen solubility parameters (HSP) distance of HPMC with various alcohols. The
4
dashed line represents the interaction radius of HPMC. (b) The change in Hansen solubility
5
6 parameters (HSP) distance of HPMC with various alcohols upon the addition of water. The
7 dashed line represents the interaction radius of HPMC.
8 The addition of water resulted in a decrease in the HSP distance in case of higher alcohols as can
9
10 be inferred from Figure 9b. Thus, the polymer is expected to be more soluble in alcohols
11
12
13
containing water.
14
15
16 4.8. Molecular Dynamics Simulations
17
18 The statistical analysis of the MD trajectories produced in this study were based on calculations
19
20
21 of the radius of gyration, Rg; a quantity commonly used in polymer physics to evaluate the
22
23 structure and dimensions of chain-like materials. For a molecular system containing N particles,
24
25 at any step of the trajectory, Rg is defined as:
26
27 ∑( '
  ! "!#$% &
28  ≡ (
∑ 
(4)
29
30
31 Average values of Rg along production trajectories are useful. Meaningful details emerged by
32
33 performing a full statistical analysis on Rg of the polymer. The summary is presented on Figure
34
35
10. In all cases, the lowest Rg averages are found at compositions that mix half alcohol and half
36
37
38 DCM. In contrast, the highest averages are found in pure PrOH and IPA, while for EtOH, the
39
40 system peaks at an alcohol proportion of 70%. The effect of increasing the proportion of the
41
42 alcohol on the conformation of the polymer is more pronounced in the case of IPA, than in the
43
44
45 case of PrOH, and mixed results are found for EtOH. Roughly, Rg tends to be larger when the
46
47 proportion of alcohol in the solvent is greater, indicating that a more extended conformation of
48
49 HPMC may be favored in the presence of alcohols.
50
51
52
Sampling can be challenging in polymer simulations. Therefore, the barriers of the polymer
53
54 backbone torsions were explored to estimate the height of relevant features of the conformational
55
56 energy surface of the polymer. The torsional angle C2-C1-O1-C4 was rigidly scanned to explore
57
58 33
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 34 of 48

1
2
3 the conformational freedom of the backbone (see Supplementary Information). These torsions
4
5
6 are expected to be a more significant restraint than those involving the side-chains. This is
7
8 mainly because the size of the side-chain substituents is small compared to the full polymer
9
10 chain. On average, the estimated energy barrier for the backbone torsion of the substituted-
11
12
13
HPMC model is approximately 7 kcal/mol. Considering that the temperature of simulation is
14
15 around 300 K, such torsions are moderately weak and may have limited impact in
16
17 conformational sampling.
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36 Figure 10. Radius of gyration, ⟨Rg⟩, for HPMC in the presence of different alcohol:DCM
37 compositions during the 20 ns NPT production stage. The average, range, standard deviation,
38 and quartiles are shown.
39
40 To analyze the effect of water in the solvent, a comparison was carried out between simulations
41
42
43 of four polymer chains with and without water. The results are summarized in Figure 11. Water
44
45 appears to have a different effect on the Rg of HPMC depending on the type of alcohol present in
46
47 the solvent. The polymer clearly expands in PrOH, while the Rg shows a limited increase in
48
49
EtOH, and it remains almost invariant in the case of IPA.
50
51
52
53
54
55
56
57
58 34
59
60 ACS Paragon Plus Environment
Page 35 of 48 Molecular Pharmaceutics

1
2
3
4
5
6
7
8
9
10
11
12
13 Figure 11. Average radius of gyration, ⟨Rg⟩, for four chains of HPMC in the presence (cyan
14 color) and absence (brown color) of water in (a) EtOH, (b) PrOH, and (c) IPA. The black
15 horizontal line corresponds to the arithmetic mean; the candlesticks are centered on the median
16 and represent the standard deviation; and the whisker bars show the lower and higher quartiles.
17
18 In addition, radial distribution functions (RDF) were determined between all the atom types of
19
20
21 the polymer chains and all the atom types of the solvent molecules. The RDF represents the
22
23 probability of finding two sets of atoms at a certain distance. Figure 12 shows the RDF for
24
25 HPMC-alcohol and HPMC-DCM for different alcohol-DCM mixtures. There are two solvation
26
27
28 layers around the HPMC polymer chain, one at approximately 0.55 nm and one at 0.95 nm.
29
30 Interestingly, for 70:30 % alcohol:DCM mixtures, the probability of finding DCM or alcohol
31
32 near the polymer chain is the same (blue lines). However, when the proportion of DCM increases
33
34
to 50%, the probability of finding DCM close to the polymer is reduced, and the probability of
35
36
37 finding alcohol near the polymer is increased (red lines).
38
39 Figure 13 shows the RDF between HPMC and solvent for systems containing 4 polymer chains
40
41 in the presence of 55:30:15 % alcohol:DCM:water mixtures. The addition of water to the
42
43
44 systems causes the creation of two water shells surrounding the polymer chain, one at
45
46 approximately 0.41 nm and one at 0.78 nm (black line). The probability of finding water close to
47
48 the polymer is higher than the probability of finding alcohol or DCM molecules. Systems with
49
50
51 PrOH and IPA alcohols exhibited a higher probability of finding water close to the polymer than
52
53 the EtOH system.
54
55
56
57
58 35
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 36 of 48

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 Figure 12. Radial Distribution Function for HPMC with the solvent in the presence of different
38 alcohol-DCM mixtures.
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 36
59
60 ACS Paragon Plus Environment
Page 37 of 48 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 Figure 13. Radial Distribution Function for HPMC with the solvent for systems with four
37 polymer chains in the presence of 55:30:15 % alcohol:DCM:water.
38
39
5. Discussion
40
41
42 The ITZ-HPMC system has been reported to undergo phase separation when spin coated and
43
12, 13
44 spray dried from a DCM:MeOH solvent mixture . A recent comprehensive investigation of
45
46 this system revealed that phase separation was driven by the presence of water, either acquired
47
48
49 from the environment or present in the solvent system. Herein, we have further studied the role
50
51 of water, expanding the study to evaluate solvent characteristics, in particular higher alcohols,
52
53 and polymer conformation on the phase behavior of the ITZ:HPMC system.
54
55
56
57
58 37
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 38 of 48

1
2
3 Solvent selection for ASD formation is often empirical and typically driven by the need to find a
4
5
6 common solvent/solvent mixture for all the ASD components. However, solvent properties can
7
8 impact the properties and microstructure of the resultant ASD differently. Findings from this
9
10 study indicate that small amounts of water, present as a contaminant in the solvent system, can
11
12
13
be a double-edged sword in terms of impact on miscibility. The presence of water in the solvent
14
15 system during ASD fabrication has been reported to lead to phase separation for a number of
16
17 systems.33, 34 ITZ-HPMC phase behavior when prepared from the lower alcohol-DCM solvent
18
19
system followed the expected trend whereby water led to phase separation and the formation of
20
21
22 discrete drug-rich domains. However, the improvement in miscibility induced by water for the
23
24 higher alcohols solvent systems, as observed herein, is a somewhat surprising observation. These
25
26 observations are of particular interest for HPMC-based ASDs, since most grades of this polymer
27
28
29 are not soluble in single component organic solvent systems, necessitating the use of
30
31 multicomponent solvent systems. To facilitate discussion of the phase behavior, the co-solvent
32
33 alcohols will be categorized into lower alcohols (MeOH and EtOH) and higher alcohols (IPA,
34
35
36
PrOH and BuOH).
37
38 When lower alcohols were chosen as co-solvents, dry conditions were necessary to ensure that
39
40 ITZ and HPMC intimately mixed in the ASD. When the spin coating process was carried out at
41
42
21 °C and 50% RH, both solvent systems led to the formation of phase separated domains, which
43
44
45 decreased in size with an increase in evaporation rate (Figure 2). When the humidity in the
46
47 environment was lowered to 20% RH, the ETOH co-solvent system resulted in a miscible ASD
48
49 while the DCM:MeOH solvent system resulted in a phase separated ASD. For the latter system,
50
51
52 only after reduction of the ambient relative humidity to 7%, could a miscible ASD be
53
54 produced.13 It is well established that the absorption of water is favored at a higher humidity due
55
56
57
58 38
59
60 ACS Paragon Plus Environment
Page 39 of 48 Molecular Pharmaceutics

1
2
3 to the higher partial pressure of water in the atmosphere.35 Thus, it is expected that less water
4
5
6 will be absorbed during spin coating at 20% RH. For EtOH, at 20% RH, the equilibrium water
7
8 content is less than 3% by volume, which is close to, but below the azeotropic composition of the
9
10 EtOH:H2O binary mixture, which is 95:5 % EtOH:H2O by volume.29 Given that the system is
11
12
13
EtOH-rich relative to the azeotropic composition, evaporation will lead to residual liquid that is
14
15 rich in EtOH and concentration of H2O will not occur in the solution. Of course, this analysis
16
17 does not consider the impact of either DCM or the dissolved solutes on the azeotropic
18
19
composition, and these are unknown factors. In contrast, MeOH does not form an azeotropic
20
21
22 system with water. Hence, MeOH is anticipated to evaporate faster than water, leading to an
23
24 increase in the H2O concentration in the residual solution. The presence of as little as 1% water
25
26 in the DCM:MeOH stock solution has been reported to induce phase separation in spray dried
27
28
29 ITZ:HPMC solid dispersions due to the poor aqueous solubility of ITZ.13 Thus, it has to be
30
31 prepared at very low humidity (7% RH) to ensure that the water content in the solvent system is
32
33 minimized and the resultant ASD is miscible.
34
35
36
Support for these suppositions is provided by the solvent evaporation and evaporation rate
37
38 profiles (Figure 6). For the solvent mixtures, without any dissolved materials, the inclusion of
39
40 4% water substantially reduced the evaporation rate of the DCM:MeOH mixture. In contrast, the
41
42
inclusion of 4% water did not impact the evaporation of DCM:EtOH solvent mixture (Figures 6a
43
44
45 and b). To summarize, both EtOH and MeOH are hygroscopic solvents that are expected to
46
47 acquire water from the atmosphere during spin coating. For EtOH, which forms an azeotrope
48
49 with H2O, during solvent evaporation, high initial water contents (>5% by volume) are expected
50
51
52 to lead to water concentration, while low initial water contents will result in EtOH concentration.
53
54 For MeOH systems, H2O concentration is anticipated during evaporation. Due to the low
55
56
57
58 39
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 40 of 48

1
2
3 aqueous solubility of ITZ, phase separation of the drug to form discrete domains is anticipated to
4
5
6 occur when water concentration occurs during solvent evaporation, and when the evaporation
7
8 rate is slow. Using this framework, our experimental observations can be largely rationalized.
9
10 Higher alcohols and H2O impacted the phase behavior of the drug-polymer films differently. IPA
11
12
13
displayed a unique behavior when used as a co-solvent with DCM. At 20% RH, the spin coated
14
15 ITZ:HPMC system was found to display a network-like corrugated structure at all of the
16
17 investigated temperatures (21 - 50 °C). The fact that the high features in the AFM topographical
18
19
images are HPMC-rich (Figure 5), coupled with the aggregation of HPMC in the IPA-containing
20
21
22 solvent system both in solution state (Figures 7a and b) or in the deposited film (Figure 8),
23
24 suggests that the observed phase separation at 20% RH is due to the incompatibility of the
25
26 polymer with the solvent system. At higher relative humidity and lower temperatures, discrete,
27
28
29 spherical domains embedded in an inhomogeneous background are observed. Not only is the
30
31 polymer poorly soluble in IPA, but ITZ also shows the lowest solubility in this alcohol (Figure
32
33 SI.2a and b). The absorption of water will further reduce the drug solubility, hence the discrete
34
35
36
domains are likely to consist of a drug-rich phase. Thus, for IPA co-solvent systems, the phase
37
38 behavior is extremely complex with both low drug and polymer solubility coming into play,
39
40 depending on the water content of the system.
41
42
PrOH and BuOH, impacted ITZ:HPMC phase behavior similarly in the co-solvent system. Both
43
44
45 solvents can form an azeotrope with water at relatively high contents of water (71.7 and 55.5%
46
47 PrOH and BuOH with water, respectively).29 Hence during evaporation, the system will
48
49 concentrate with respect to the alcohol and not H2O. This is likely why the presence of water did
50
51
52 not markedly impact the evaporation time (Figure 5). Interestingly, the presence of water was
53
54 necessary to ensure solubilization of HPMC in these higher alcohols systems, based on the
55
56
57
58 40
59
60 ACS Paragon Plus Environment
Page 41 of 48 Molecular Pharmaceutics

1
2
3 formation of phase separated films when produced at low RH, as well as the polymer
4
5
6 aggregation during evaporation under dry conditions as illustrated by DLS results (Figure 7) and
7
8 film texture investigated using AFM (Figure 8). Hansen sphere theory suggests that the longer
9
10 the alcohol chain, the lower the miscibility of the polymer in that alcohol, and the miscibility also
11
12
13
decreases with branching as in the case of IPA (Figure 9a). The analysis also indicates that the
14
15 addition of water increases the miscibility of the polymer in the higher alcohols (Figure 9b). The
16
17 impact of higher alcohol viscosity on the phase behavior of this system also needs to be
18
19
considered. The viscosity of alkyl alcohols increases with an increase in alkyl chain length with
20
21 36 37
22 the following literature values reported at 20°C: MeOH= 0.55 cP , EtOH= 1.19 cP , IPA=
23
36 36 36
24 2.27 cP , PrOH= 2.2 cP and BuOH= 2.95 cP . It is generally accepted that an increase in
25
26 viscosity results in a decrease in molecular diffusivity which in turn decreases the nucleation rate
27
28
29 of a new phase.38 Thus, higher alcohols are expected to reduce the diffusion of solute molecules
30
31 to form drug-rich domains.
32
33 Overall observations are summarized in Figure 14. Although all initial solutions were visually
34
35
36
transparent suggesting complete solubility of solid components in the solvent mixture, the
37
38 polymer was found to exist as aggregateswhen higher alcohols were used as a co-solvent
39
40 (Figures 7a and b). This solvophobic behavior of HPMC in higher alcohols is responsible for the
41
42
production of irregular HPMC films under dry conditions (Figure 8). Luo and Eisenberg have
43
44
45 reported that the aggregate size for a block copolymer depends on solvent composition for a
46
47 complex organic solvent mixture.39
48
49
50
51
52
53
54
55
56
57
58 41
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 42 of 48

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 Figure 14. Schematic representation of the impact of water on the phase behavior of ITZ-HPMC
36
solid dispersion. Green checks indicate the compatibility of the component with the solvent
37
38 while red crosses indicate incompatibility of the component with the solvent.
39
40
41 Our data suggests that HPMC needs to exist as discrete molecules in the extended conformation
42
43 in the solvent system in order to produce an intimately mixed ASD. In addition, the discrete
44
45
46 domains and spinodal decomposition seem to be induced by incompatibility of the drug with
47
48 water present in the solvent system, while the network-like corrugations are likely induced by
49
50 solvent incompatibility with HPMC. It is anticipated that the presence of water enables the
51
52
53
polymer to remain as a single chain in the preferred extended conformation in the higher
54
55 alcohols. It has been reported that extended polymer conformations are important to promote
56
57
58 42
59
60 ACS Paragon Plus Environment
Page 43 of 48 Molecular Pharmaceutics

1
2
3 interaction with drug molecules and prevent drug crystallization.40 The impact of solvent on the
4
5
6 preferred conformation of hydrophilic polymers such as polyethylene glycols (PEG) has been
7
8 investigated,41 whereby three molecules of water were found to be required for hydration of each
9
10 ethylene oxide unit in a PEG chain. The influence of hydration shell structure on the
11
12
13
conformational stability and function of hydrophilic polymers and biomolecules including
14
15 proteins and DNA has been extensively investigated in the literature. 42 In the present study, MD
16
17 simulations were useful to determine how variations in the solvent composition influence
18
19
variables such as the Rg of HPMC, and the RDF between polymer and solvent atoms. The Rg
20
21
22 values were susceptible to the composition of the solvent mixture, with the lowest Rg values
23
24 observed for systems with a 50:50 % DCM:alcohol composition (Figure 10). The addition of
25
26 more DCM to the mixture caused a lower probability of finding DCM close to the HPMC chain
27
28
29 and instead, an increased probability of finding alcohol (Figure 12). This is an important
30
31 phenomenon observed for all the alcohol-DCM systems, which may imply that during the
32
33 evaporation process of the alcohol-DCM mixture, the solvation shell around the polymer will
34
35
36
constantly vary depending on the concentration of both solvents. These changes in the solvent
37
38 composition can subsequently cause variation in the Rg of the polymer. Therefore, the miscibility
39
40 results shown in Figure 8 will not only depend on the polymer conformation in the initial 50:50
41
42
mixture, but also on how the polymer conformation changes due to variations in solvent
43
44
45 composition during the evaporation process.
46
47 Additionally, to determine the effect of the relative humidity in the properties of the HPMC
48
49 chain, simulations were conducted for 55:30:15 % alcohol:DCM:water mixtures. The RDF
50
51
52 between HPMC and solvent for these systems (Figure 13) shows that the first layer of solvent
53
54 surrounding the polymer is predominantly formed by water molecules. Thus, there is a higher
55
56
57
58 43
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 44 of 48

1
2
3 probability of finding water than DCM or alcohol within 1 nm of the polymer. This suggests that
4
5
6 water will solvate the HPMC polymer chains for DCM-alcohol mixtures that tend to absorb
7
8 water under high relative humidity. That solvation by water molecules appeared to increase the
9
10 Rg for HPMC in the presence of PrOH-DCM-water mixtures (Figure 11), is in agreement with
11
12
13
the results presented in Figure 8 showing the change in film microstructure. These observations
14
15 around the presence of a water shell around the polymer chain could have important implications
16
17 for the conformation of the polymer chain as a function of solvent composition, and also in the
18
19
possible phase separation of polymer-drug mixtures.
20
21
22 Conclusions
23
24 Evaluation of the impact of solvent composition and evaporation rate on the phase behavior of
25
26 spin coated ASDs revealed that water can be a double edged sword in terms of drug-polymer
27
28
29 miscibility for a the ITZ:HPMC system. Based on the physicochemical properties of the drug
30
31 and polymer, in particular their solubility in the dry and water-containing solvent system, the
32
33 phase behavior of a specific system can vary even with slight changes in solvent composition.
34
35
36
Domain sizes of phase separated regions were also found to decrease with increasing
37
38 temperature. This study highlights the complex interplay of component properties and formation
39
40 conditions on the resultant ASD microstructure and ultimately, these observations may help in
41
42
the rational design of ASDs produced via solvent evaporation.
43
44
45 ACKNOWLEDGMENTS
46
47 We are grateful to the National Institute for Pharmaceutical Technology and Education (NIPTE)
48
49 and the U.S. Food and Drug Administration (FDA) for providing funds for this research. This
50
51
52 study was funded by the FDA Grant to NIPTE titled "The Critical Path Manufacturing Sector
53
54 Research Initiative (U01)"; Grant# 5U01FD004275. The Pharmaceutical Research and
55
56
57
58 44
59
60 ACS Paragon Plus Environment
Page 45 of 48 Molecular Pharmaceutics

1
2
3 Manufacturers of America (PhRMA) Foundation is acknowledged for a post-doctoral fellowship
4
5
6 granted to NM. We also express our gratitude for funding provided by a Graduate Student
7
8 Fellowship Award from the American Association of Pharmaceutical Scientists (AAPS), the
9
10 McKeehan Graduate Fellowship in Pharmacy, and the Migliaccio/Pfizer Graduate Fellowship in
11
12
13
Pharmaceutical Sciences, awarded to LM. This research was supported in part through
14
15 computational resources provided by Information Technology at Purdue University.
16
17
18
19
Supporting Information
20
21
22 Fluorescence images of spin-coated drug-polymer films; clear and cloudpoint data; X-ray
23
24 diffraction data of precipitates collected from cloudpoint experiments; DSC thermograms of
25
26 precipitates collected from cloudpoint experiments; torsional energy profiles for HPMC dimers
27
28
29 used in modeling; range of torsional energies for various HPMC dimers.
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 45
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 46 of 48

1
2
3 References
4
5
(1) Simonelli, A. P. M., S. C.; Higuchi, W. I., Dissolution rates of high energy sulfathiazole-povidone
6
7
coprecipitates II: Characterization of form of drug controlling its dissolution rate via solubility studies. J.
8 Pharm. Sci. 1976, 65, (3), 355−361.
9 (2) Rumondor, A. C. F.; Stanford, L. A.; Taylor, L. S., Effects of Polymer Type and Storage Relative
10 Humidity on the Kinetics of Felodipine Crystallization from Amorphous Solid Dispersions. Pharm. Res.
11 2009, 26, (12), 2599-2606.
12 (3) Caron, V.; Hu, Y.; Tajber, L.; Erxleben, A.; Corrigan, O. I.; McArdle, P.; Healy, A. M., Amorphous
13 Solid Dispersions of Sulfonamide/Soluplus (R) and Sulfonamide/PVP Prepared by Ball Milling. AAPS
14 Pharm. Sci. Tech. 2013, 14, (1), 464-474.
15
(4) Sun, Y.; Tao, J.; Zhang, G. G. Z.; Yu, L. A., Solubilities of Crystalline Drugs in Polymers: An
16
17 Improved Analytical Method and Comparison of Solubilities of Indomethacin and Nifedipine in PVP,
18 PVP/VA, and PVAc. J Pharm Sci 2010, 99, (9), 4023-4031.
19 (5) Serajuddin, A. T. M., Solid dispersion of poorly water-soluble drugs: Early promises, subsequent
20 problems, and recent breakthroughs. J Pharm Sci 1999, 88, (10), 1058-1066.
21 (6) Zhang, J. X.; Bunker, M.; Parker, A.; Madden-Smith, C. E.; Patel, N.; Roberts, C. J., The Stability of
22 Solid Dispersions of Felodipine in Polyvinylpyrrolidone Characterized by Nanothermal Analysis. Int. J.
23 Pharm. 2011, 414, (1-2), 210-217.
24
(7) Bank, M.; Leffingwell, J.; Thies, C., Influence of Solvent Upon Compatibility of Polystyrene and
25
26 Poly(Vinyl Methyl Ether). Macromolecules 1971, 4, (1), 43-46.
27 (8) Paudel, A.; Van den Mooter, G., Influence of Solvent Composition on the Miscibility and Physical
28 Stability of Naproxen/PVP K 25 Solid Dispersions Prepared by Cosolvent Spray-Drying. Pharm. Res. 2012,
29 29, (1), 251-270.
30 (9) Ansari, M. T.; Sunderland, V. B., Solid dispersions of dihydroartemisinin in polyvinylpyrrolidone.
31 Arch Pharm Res 2008, 31, (3), 390-398.
32 (10) Al-Obaidi, H.; Brocchini, S.; Buckton, G., Anomalous properties of spray dried solid dispersions. J
33
Pharm Sci 2009, 98, (12), 4724-37.
34
35
(11) Wulsten, E.; Kiekens, F.; van Dycke, F.; Voorspoels, J.; Lee, G., Levitated single-droplet drying:
36 Case study with itraconazole dried in binary organic solvent mixtures. Int J Pharm 2009, 378, (1), 116-
37 121.
38 (12) Purohit, H. S.; Taylor, L. S., Miscibility of Itraconazole-Hydroxypropyl Methylcellulose Blends:
39 Insights with High Resolution Analytical Methodologies. Mol Pharmaceut 2015, 12, (12), 4542-4553.
40 (13) Mugheirbi, N. A.; Marsac, P. J.; Taylor, L. S., Insights into Water-Induced Phase Separation in
41 Itraconazole-Hydroxypropylmethyl Cellulose Spin Coated and Spray Dried Dispersions. Mol Pharmaceut
42 2017, 14, (12), 4387-4402.
43
(14) Ricarte, R. G.; Lodge, T. P.; Hillmyer, M. A., Nanoscale Concentration Quantification of
44
45 Pharmaceutical Actives in Amorphous Polymer Matrices by Electron Energy-Loss Spectroscopy.
46 Langmuir 2016, 32, (29), 7411-7419.
47 (15) Ricarte, R. G.; Lodge, T. P.; Hillmyer, M. A., Detection of Pharmaceutical Drug Crystallites in Solid
48 Dispersions by Transmission Electron Microscopy. Mol Pharmaceut 2015, 12, (3), 983-990.
49 (16) Mugheirbi, N. A.; O'Connell, P.; Serrano, D. R.; Healy, A. M.; Taylor, L. S.; Tajber, L., A
50 Comparative Study on the Performance of Inert and Functionalized Spheres Coated with Solid
51 Dispersions Made of Two Structurally Related Antifungal Drugs. Mol Pharmaceut 2017, 14, (11), 3718-
52
3728.
53
54
(17) Hansen. C. M, The Three Dimensional Solubility Parameter and Solvent Diffusion Coefficient Their
55 Importance In Surface Coating Formulation ed.; Danish Technical Press Copenhagen, 1967.
56
57
58 46
59
60 ACS Paragon Plus Environment
Page 47 of 48 Molecular Pharmaceutics

1
2
3 (18) Archer, W. L., Hansen Solubility Parameters for Selected Cellulose Ether Derivatives and Their
4
Use in the Pharmaceutical-Industry. Drug Dev Ind Pharm 1992, 18, (5), 599-616.
5
(19) Sorensen, P., Solubility Parameter Concept in Formulation of Liquid Inks. J Oil Colour Chem As
6
7 1967, 50, (3), 226-&.
8 (20) https://www.hansen-solubility.com/downloads.php.
9 (21) Jiang, X.; Wang, Y.; Li, M., Selecting water-alcohol mixed solvent for synthesis of polydopamine
10 nano-spheres using solubility parameter. Scientific Reports 2014, 4, 6070.
11 (22) Van der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A. E.; Berendsen, H. J. C., GROMACS:
12 Fast, flexible, and free. J Comput Chem 2005, 26, (16), 1701-1718.
13 (23) Brooks, B. R.; Bruccoleri, R. E.; Olafson, B. D.; States, D. J.; Swaminathan, S.; Karplus, M.,
14
Charmm - a Program for Macromolecular Energy, Minimization, and Dynamics Calculations. J Comput
15
16 Chem 1983, 4, (2), 187-217.
17 (24) MacKerell, A. D.; Banavali, N.; Foloppe, N., Development and current status of the CHARMM
18 force field for nucleic acids. Biopolymers 2001, 56, (4), 257-265.
19 (25) http://virtualchemistry.org/
20 (26) Zoete, V.; Cuendet, M. A.; Grosdidier, A.; Michielin, O., SwissParam: A Fast Force Field
21 Generation Tool for Small Organic Molecules. J Comput Chem 2011, 32, (11), 2359-2368.
22 (27) Essmann, U.; Perera, L.; Berkowitz, M. L.; Darden, T.; Lee, H.; Pedersen, L. G., A Smooth Particle
23
Mesh Ewald Method. J Chem Phys 1995, 103, (19), 8577-8593.
24
25
(28) Froimowitz, M., Hyperchem(Tm) - a Software Package for Computational Chemistry and
26 Molecular Modeling. Biotechniques 1993, 14, (6), 1010-1013.
27 (29) Carper, J., The CRC Handbook of Chemistry and Physics. Libr J 1999, 124, (10), 192-+.
28 (30) Mugheirbi, N. A.; Tajber, L., Crystal Habits of Itraconazole Microcrystals: Unusual Isomorphic
29 Intergrowths Induced via Tuning Recrystallization Conditions. Molecular Pharmaceut 2015, 12, (9), 3468-
30 3478.
31 (31) Mugheirbi, N. A.; Paluch, K. J.; Tajber, L., Heat induced evaporative antisolvent
32 nanoprecipitation (HIEAN) of itraconazole. Int J Pharm 2014, 471, (1-2), 400-11.
33
(32) Milliman, H. W.; Boris, D.; Schiraldi, D. A., Experimental Determination of Hansen Solubility
34
35 Parameters for Select POSS and Polymer Compounds as a Guide to POSS-Polymer Interaction Potentials.
36 Macromolecules 2012, 45, (4), 1931-1936.
37 (33) Saboo, S.; Taylor, L. S., Water-induced phase separation of miconazole-poly (vinylpyrrolidone-
38 co-vinyl acetate) amorphous solid dispersions: Insights with confocal fluorescence microscopy. Int J
39 Pharm 2017, 529, (1-2), 654-666.
40 (34) Rumondor, A. C. F.; Wikstrom, H.; Van Eerdenbrugh, B.; Taylor, L. S., Understanding the
41 Tendency of Amorphous Solid Dispersions to Undergo Amorphous-Amorphous Phase Separation in the
42
Presence of Absorbed Moisture. AAPS Pharmscitech 2011, 12, (4), 1209-1219.
43
44
(35) De Vrieze, S.; Van Camp, T.; Nelvig, A.; Hagstrom, B.; Westbroek, P.; De Clerck, K., The effect of
45 temperature and humidity on electrospinning. J Mater Sci 2009, 44, (5), 1357-1362.
46 (36) https://www.fishersci.com/us/en/brands/f/fisher-chemical.html.html
47 (37) Kadlec, P.; Henke, S.; Bubnik, Z., Properties of ethanol and ethanol-water solutions - Tables and
48 Equations. Zuckerindustrie 2010, 135, (10), 607-613.
49 (38) Warren, D. B.; Benameur, H.; Porter, C. J.; Pouton, C. W., Using polymeric precipitation inhibitors
50 to improve the absorption of poorly water-soluble drugs: A mechanistic basis for utility. J Drug Targeting
51
2010, 18, (10), 704-731.
52
(39) Luo, L. B.; Eisenberg, A., Thermodynamic size control of block copolymer vesicles in solution.
53
54 Langmuir 2001, 17, (22), 6804-6811.
55
56
57
58 47
59
60 ACS Paragon Plus Environment
Molecular Pharmaceutics Page 48 of 48

1
2
3 (40) Mosquera-Giraldo, L. I.; Borca, C. H.; Meng, X. T.; Edgar, K. J.; Slipchenko, L. V.; Taylor, L. S.,
4
Mechanistic Design of Chemically Diverse Polymers with Applications in Oral Drug Delivery.
5
6
Biomacromolecules 2016, 17, (11), 3659-3671.
7 (41) Liu, K. J.; Parsons, J. L., Solvent Effects on Preferred Conformation of Poly(Ethylene Glycols).
8 Macromolecules 1969, 2, (5), 529-&.
9 (42) Mochizuki, K.; Ben-Amotz, D., Hydration-Shell Transformation of Thermosensitive Aqueous
10 Polymers. J Phys Chem Lett 2017, 8, (7), 1360-1364.
11
12
13
14
15
16 TOC
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 48
59
60 ACS Paragon Plus Environment

You might also like