Qian 2010

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

COMMENTARY

Drug–Polymer Solubility and Miscibility: Stability


Consideration and Practical Challenges in Amorphous
Solid Dispersion Development
FENG QIAN, JUN HUANG, MUNIR A. HUSSAIN
Bristol-MyersSquibb Co., Biopharmaceutics R&D, One Squibb Dr., Bldg. 105/Room 2474, New Brunswick, New Jersey 08903

Received 15 September 2009; revised 1 December 2009; accepted 2 December 2009


Published online in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/jps.22074

ABSTRACT: Drug–polymer solid dispersion has been demonstrated as a feasible approach to


formulate poorly water-soluble drugs in the amorphous form, for the enhancement of dissolution
rate and bioperformance. The solubility (for crystalline drug) and miscibility (for amorphous
drug) in the polymer are directly related to the stabilization of amorphous drug against
crystallization. Therefore, it is important for pharmaceutical scientists to rationally assess
solubility and miscibility in order to select the optimal formulation (e.g., polymer type, drug
loading, etc.) and recommend storage conditions, with respect to maximizing the physical
stability. This commentary attempts to discuss the concepts and implications of the drug–
polymer solubility and miscibility on the stabilization of solid dispersions, review recent
literatures, and propose some practical strategies for the evaluation and development of such
systems utilizing a working diagram. ß 2010 Wiley-Liss, Inc. and the American Pharmacists Associa-
tion J Pharm Sci 99:2941–2947, 2010
Keywords: amorphous solid dispersion; solubility; miscibility; physical stability; glass
transition temperature; drug loading

INTRODUCTION dispersions, wherein an amorphous drug–polymer


composite is generated by methods such as spray
Development of amorphous drugs has been an drying and hot-melt extrusion, has become one of the
inspiring topic in pharmaceutical research for both most promising and viable amorphous formulation
industrial and academic scientists. With the fact that approaches.7,8 Compared with other noncrystalline
many compounds from current discovery programs drug delivery approaches such as cosolvency or self-
are poorly water-soluble, amorphous drugs are emulsifying drug delivery systems (SEDDS),9,10
advantageous over their crystalline counterparts amorphous solid dispersions are more amenable to
with higher solubility, faster dissolution rate, and be developed into preferred solid dosage forms as the
enhanced oral bioavailability.1,2 On the other hand, final drug product, which often leads to highly
amorphous solids are physically unstable relative to desirable advantages over liquid or semisolid for-
the crystalline state and the stability of amorphous mulations including lower manufacturing cost, smal-
drugs against crystallization is critical for pharma- ler pill burden, improved physical and chemical
ceutical development because crystallization negates stability, possibility of combination dosage forms,
the advantages of amorphous solids.3,4 etc. For clarification, ‘‘solid dispersion’’ discussed in
Since initially introduced by Sekiguchi et al.,5,6 this commentary refers to the most widely investi-
solid dispersions have been drawing continuous gated amorphous solid dispersion, wherein a crystal-
interests for the enhancement of drug bioavailability. lizable, small molecule drug disperses in an
Among a variety of solid dispersions, amorphous amorphous polymer matrix. Systems containing
semicrystalline polymers like PEG11–13 or Pluro-
nic114 involve other scientific considerations and will
Correspondence to: Munir A. Hussain (Telephone: 732-227-
3272; Fax: 732-227-3752; E-mail: munir.hussain@bms.com) not be included in this discussion.
Journal of Pharmaceutical Sciences, Vol. 99, 2941–2947 (2010) Due to the lack of adequate understanding of the
ß 2010 Wiley-Liss, Inc. and the American Pharmacists Association nonequilibrium glassy state and other practical

JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 99, NO. 7, JULY 2010 2941


2942 QIAN, HUANG AND HUSSAIN

complexities,15 the mechanisms as to why solid partially miscible and the A-rich two-component
dispersions can stabilize amorphous drugs are not phase coexists with the B-rich phase and their
yet fully understood besides several commonly equilibrium compositions are determined by the
accepted viewpoints. For example, the addition of a two miscibility values on the curve. Under the scope
high-Tg (glass transition temperature) polymer ele- of this commentary, the compositional miscibility of
vates the Tg of the amorphous system and therefore drug in polymer is practically more relevant than the
may reduce its molecular mobility required for miscibility of the polymer in the drug in the sense that
crystallization at certain storage temperatures.3 polymer is more likely to be the dominant phase in
Thermodynamically the drug has a lower chemical these systems for typical small molecule drug–
potential when mixed with a polymer, resulting in the polymer solid dispersions developed in pharmaceu-
change of crystallization driving force. It is also tical industry. In other words, our discussions will
generally accepted that drug–polymer intermolecular focus on the temperature/composition range within
interactions are important for the stabilization of which the drug–polymer miscibility may be regarded
the solid dispersion.16–18 Underneath all the above as the ‘‘solubility’’ of the amorphous drug in the
considerations there is an important assumption that polymer.
the two components (drug and polymer) are mixed The concept of miscibility is well understood in
homogeneously at the molecular level. This assump- polymer physics for polymer blends,20 in which each
tion is directly related to the solubility (for the polymer component is usually in stable amorphous
crystalline drug) and/or miscibility (for the amor- state. The major complexity around the miscibility of
phous drug) in the polymer matrix, two terms that a small molecule drug in a solid dispersion is that the
appear straightforward while have sometimes been amorphous drug is usually meta-stable relative to the
used with confusion. This commentary briefly dis- crystalline state and inclines to crystallize. Thus, the
cusses the drug–polymer solubility and miscibility system would eventually reach equilibrium with
with regard to the concepts and their implications to regard to the crystalline drug and the equilibrium
the stabilization of solid dispersions (e.g., polymer composition of the mixture would be the solubility of
selection, stability assessment, etc.), within the scope the crystalline drug in the polymer. The measurable
of industrial formulation strategies. Literature miscibility in the case of drug–polymer dispersion is
review of the current progress on the determina- therefore associated with a meta-stable equilibrium
tion/estimation of the solubility and miscibility will different from what was originally defined in polymer
also be discussed. science and requires that the drug stay in the super-
cooled liquid (a liquid at temperature below the
crystalline melting point Tm and above Tg) state
DRUG–POLYMER SOLUBILITY AND without crystallization (a ‘‘stable’’ meta-stable state,
MISCIBILITY: CONCEPTS AND IMPLICATIONS similar to the situation of coexisting polymorphs21)
within the experimental time frame. Similar to
In regular small-molecule solutions, solubility is solubility, miscibility at temperatures close to and
defined as an equilibrium thermodynamic parameter, below Tg is only apparent and involves the kinetics of
at which the chemical potential of the solute in the phase separation and structural relaxation. The
solid phase is the same as that in the liquid (solution) theoretical equilibrium miscibility practically may
phase. This definition can be extended for polymer only be estimated from extrapolation and modeling.
solvents assuming equilibrium can always be reached In the study of polymer blends of two stable
when the temperature of measurement is well-above amorphous components, the miscibility behavior is
Tg of the system. When the temperature is close to or typically described by the well-known Flory–Huggins
below Tg, the system becomes too viscous to maintain theory,22,23 a lattice-based, statistical mechanics
equilibrium and the solubility determined experi- model where the free energy of mixing is broken into
mentally (if possible) lacks a clear physical meaning two parts: an entropy part that always favors mixing
and can only be defined as ‘‘apparent’’ to be dis- and an enthalpy part that can either facilitate or
tinguished from the equilibrium solubility. The prevent mixing, depending on the nature and
equilibrium solubility at low temperatures may be intensity of the interaction between the two compo-
estimated by extrapolation or model prediction. nents. At a given temperature, complete, partial, or
Unlike the solubility curves we usually encounter, zero miscibility can be obtained with attractive
a common miscibility curve in the phase diagram of (Flory–Huggins interaction parameter x  0), weak
an A–B binary mixture has the ‘‘inverted U’’ shape as repulsive (x > 0), or strong repulsive (x  0) interac-
exemplified by the phenol–water phase diagram.19 tions, respectively. Here miscibility is the equilibrium
Above the upper critical temperature Tc, A and B are composition of the two components, above which the
completely miscible with each other; while below the free energy of mixing is greater than zero and phase
upper critical temperature, the system becomes separation is thermodynamically favorable. The

JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 99, NO. 7, JULY 2010 DOI 10.1002/jps
DRUG–POLYMER SOLUBILITY AND MISCIBILITY 2943

application of Flory–Huggins theory to the drug– amorphous drug–polymer miscibility, divide this
polymer dispersions has been shown useful while diagram into six different zones (labeled I–VI).
with limitations as well.24–26 Since amorphous drug has higher chemical potential
The importance of drug–polymer solubility and than its crystalline counterpart, the drug–polymer
miscibility in the development of amorphous solid miscibility (solubility of amorphous drug) is always
dispersions has been recognized widely. From the higher than the crystalline drug–polymer solubility.
industrial aspect, key questions include, but not Out of these data sets, Tg values of solid disper-
limited to, what the implications of these concepts are sions, which define the boundary between the
to the physical stability of solid dispersions during equilibrium liquid (Zones II, IV, and VI) and glass
the formulation, process development, and product (I, III, and V) can be obtained experimentally by
shelf-life assessment, and how we could apply the thermal analysis without ambiguity. Many times Tg
information to facilitate decision making and prior- of a binary solid dispersion can also be estimated with
itize resource allocation. It is helpful to address these reasonable accuracy by weight average theories like
questions through a comprehensive examination of a Fox or Gordon–Taylor equations with small positive
solid dispersion system, using a working diagram or negative deviations due to drug–polymer interac-
defined by temperature and drug loading in polymer tions.24,27,28
(Fig. 1) with the assumption that the system is It is generally recognized that Tg represents a
partially miscible above Tg (i.e., Tc > Tg). As stated kinetic boundary of molecular mobility. Here we
earlier, only temperature and composition range that assume that the width (temperature range) of this
has most practical interest is shown for clarification. boundary is zero and thus to the right of the Tg curve
In theory, at drug loadings above the miscibility in Figure 1 is the equilibrium liquid where structural
curve, phase separation is thermodynamically relaxation occurs rapidly and solubility and misci-
favored. At equilibrium, a drug-rich phase coexists bility can be measured at equilibrium. To the left of
with a polymer-rich phase with two distinguished Tg the Tg curve is the nonequilibrium, glassy region,
values, respectively. Practically, solid dispersions where molecular mobility is low and structural
made by industrial processes such as spray drying relaxation occurs slowly in the experimental time
or hot-melt extrusion are typically well mixed scale. Thus, equilibrium solubility and miscibility
kinetically and only show single Tg initially regard- cannot be strictly defined or experimentally mea-
less of the drug loading. For simplicity, the Tg values sured. Below Tg, the stability of the solid dispersions
shown in Figure 1 correspond to the single-phase relies heavily on the kinetics of phase separation and/
dispersion immediately after preparation. or crystallization instead of thermodynamics. In fact,
As shown in Figure 1, three data sets, Tg of the solid it has been proposed that at T < Tg  508C, the
dispersion, crystalline drug–polymer solubility, and molecular mobility can be neglected and the amor-
phous solids are stable enough over a period of years.3
The solubility curve is the boundary between the
thermodynamically stable (Zones I–II) and unstable
(Zones III–VI) regions, which defines the maximum
drug loading that can be mixed with the polymer
without worrying about crystallization or phase
separation. Zones I and II are therefore the safe
regions, within which any drug concentration and/or
temperature fluctuation of solid dispersions will not
destabilize the system. In fact, even if the drug starts
as crystalline material, the thermodynamic driving
force in Zones I and II is to dissolve the crystalline
drug in the polymer. For example, in the design of hot-
melt extrusion process, the selection of a processing
temperature at certain drug loading within Zone II
ensures thermodynamically the complete dissolution
of the drug in the polymer, assuming the dissolution
Figure 1. A hypothetical diagram consists of drug–poly-
rate is fast enough at T > Tg (hot-melt extrusion in
mer solubility, miscibility, and glass transition tempera-
Zone I is usually not feasible because of the high
tures of a solid dispersion system. A few notes: (1) the
system is partially miscible with Tc > Tg. (2) The diagram viscosity in the glassy state).
includes only temperature and composition range of main If the drug is relatively stable without crystal-
practical relevance. (3) Phase separation at drug loadings lization as a super-cooled liquid, the amorphous
higher than miscibility is ignored for simplicity when the Tg drug–polymer miscibility curve (should it be obtained
curve is drawn (see text for detail). with confidence) might represent a meta-stable

DOI 10.1002/jps JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 99, NO. 7, JULY 2010
2944 QIAN, HUANG AND HUSSAIN

boundary similar to the situation in regular solution drug loading is necessary to satisfy the dose require-
crystallization: above miscibility (Zones V and VI), ment.
spontaneous phase separation would drive the It is worth noting that Figure 1 represents a system
destabilization of solid dispersions while below where the drug is partially miscible with the polymer
miscibility but above solubility (Zones III and IV), a across the relevant temperatures. In cases where the
local fluctuation of drug concentration might be drug and the polymer are completely miscible at these
necessary to cause the destabilization. In this temperatures (e.g., indomethacin–PVP, nifedipine–
argument, the miscibility curve might shine lights PVP25,26), the miscibility curve would shift to the left
on potential strategies to enhance drug loading and and may not cross the Tg curve. Practically, the
stability of a solid dispersion, if crystallization can be miscibility line then will not appear in the working
prevented in the mixture. Above miscibility, there is diagram, while similar discussions as above still hold
no thermodynamic barrier to prevent the system from for a simplified system with four zones (I–IV).
destabilization (thus Zone VI should be avoided at
anytime) and a solid dispersion can only be stabilized
kinetically (e.g., storage at temperatures sufficiently DRUG–POLYMER SOLUBILITY AND
below Tg, Zone V). A solid dispersion below miscibility MISCIBILITY: CHALLENGES IN DETERMINATION
might be able to maintain its thermodynamically
meta-stable status both below and above Tg depend- Although drug–polymer solubility and miscibility are
ing on the crystallization behavior of the drug. important in solid dispersion development, no well-
Especially in Zone III, a solid dispersion may be established method is currently available to measure
stabilized both thermodynamically (below miscibility) them. In both of the measurements, one difficulty is
and kinetically (low molecular mobility below Tg). the high viscosity of the polymer that makes the
Although still controversial, the drug–polymer mis- system hard to reach equilibrium. On the other hand,
cibility might be an important consideration for the at temperatures near and below Tg, the system goes
physical stability of a solid dispersion, especially on continuously slow relaxation to reach equilibrium
when a homogeneous amorphous mixture can be and therefore the equilibrium solubility and misci-
obtained during processing. Table 1 summarizes the bility can only be estimated by model prediction. In
different properties of the six zones according to their this section of the commentary, we will first go
physical nature and destabilization driving force. through some literature reports on the measurement
Two concentration values, C1 and C2 as shown in of drug–polymer solubility above or close to Tg
Figure 1, are of particular interest. C1 is the drug followed by the discussion of miscibility assessment
solubility in polymer at Tg. A solid dispersion by various techniques and model application.
formulated at C1 would be thermodynamically stable So far there is no literature report for the direct
at temperatures above Tg while may be kinetically experimental measurement of the solubility of a
stable below Tg. Therefore, C1 may represent the crystalline drug in a polymer at temperatures below
highest drug loading to be adopted in a solid Tg. This is because the crystallization is extremely
dispersion that could withstand temperature fluctua- slow to achieve drug–polymer equilibrium in a glassy
tion across Tg during storage and downstream state. In fact, it was shown by Tao et al.29 that
processing, with minimal stability risk. C2, the crystalline drug–polymer equilibrium can be approxi-
miscibility of the drug in the polymer at Tg, may mately reached in the time of months even at
have a similar implication with respect to the solid temperatures well above Tg and in the presence of
dispersion formulation design, if the drug can remain crystalline seeds. In that study a DSC method was
amorphous during the time frame of formulation proposed to determine the drug–polymer solubility
processing and storage. Since C2 is higher than C1, it through direct measurement of the melting point of
may be of practical importance when extremely high the crystalline drug in the presence of the polymer

Table 1. Thermodynamic Nature and Destabilization Driving Force for Solid Dispersions in Zones I–VI of Figure 1,
Assuming All Solid Dispersions Were Homogeneously Mixed Initially

Zone Thermodynamic Nature Destabilization Driving Force

I Thermodynamically stable glass None


II Thermodynamically stable liquid None
III Supersaturated glass Crystallization of supersaturated drug
IV Supersaturated liquid Crystallization of supersaturated drug
V Supersaturated and immiscible glass Amorphous phase separation, crystallization of supersaturated drug
VI Supersaturated and immiscible liquid Amorphous phase separation, crystallization of supersaturated drug

JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 99, NO. 7, JULY 2010 DOI 10.1002/jps
DRUG–POLYMER SOLUBILITY AND MISCIBILITY 2945

(i.e., the equilibrium melting point of the mixture at estimate the drug–polymer miscibility by the deter-
this composition). This method involves sufficient mination of the interaction parameter x. The free
blending of the drug and the polymer through energy of mixing was then calculated from the Flory–
cryomilling, followed by DSC scanning with multiple Huggins equation. In their work, x was determined
slow ramping rates (18C/min or slower). The melting either experimentally from melting-point depression
end-set extrapolated to 08C/min ramping rate was data or derived from the activity of the drug in the low
recorded as the equilibrium melting temperature. molecular weight analog of the polymer. Both
The merit of this method is that it does not involve any methods, however, have limitations and need further
model prediction and can be used for most drug– verifications. The x value determined from melting-
polymer systems. The major limitation is that only point depression is applicable only within a narrow
solubility at temperatures 308C above Tg could be temperature range near the melting point and cannot
obtained with confidence, while at lower tempera- be used as a constant for lower temperature predic-
tures the system becomes too viscous to achieve tion. As observed in polymer science, x is a function of
dissolution equilibrium in the timescale of DSC both temperature and composition that can be
measurement. Prior to Tao et al.’s work, Vasantha- empirically described as below
vada et al.30,31 measured the solubility of trehalose in
dextran and PVP in the presence of moisture. The B
x¼Aþ þ CF þ DF2
solubility was determined on the basis of the Tg T
change as a function of the mixture composition and
where A, B, C, and D are constants, T and F are
moisture content. The study was actually done in a
temperature and composition, respectively.42,43
three-component system and the measurement tem-
In the other approach where x at low temperatures
perature was much more above the Tg of the system.
was calculated with the knowledge of the activity
Compared with the limited number of studies on
coefficient (g) of the drug in the polymer, g was
the solubility of crystalline drug in polymers, the
derived by solubility measurement of the drug in a
amorphous drug–polymer miscibility has been a
low molecular weight polymer analog (i.e., 1-ethyl-2-
popular topic in the literature. So far the most
pyrrolidone).26 This method estimates x at the same
common technique used to determine the mixing
temperature that the solubility was measured and
state of a solid dispersion is DSC, in which a single Tg
therefore is not limited to high temperature as in the
(intermediate of the two Tg values of the two
case of melting-point depression. The model, how-
components) indicates homogeneous mixing while
ever, requires assumptions that are not yet verified in
two Tg values indicate not. Observing the resolution
such systems (e.g., identical interaction between
limitation of the DSC method (30 nm)32 to detect
drug–polymer and drug–monomer), and is also only
microstructural phase separation, other techniques
applicable to polymers like PVP that has liquid
were reported to address the miscibility issues of solid
oligomers.
dispersions. Those include spectroscopic studies such
The g and x values determined from these studies
as IR and Raman;18 isothermal water vapor sorption
were also applied for drug–polymer solubility estima-
combined with modeling;24 spin-lattice relaxation by
tion using the solubility equation and bridging x with
solid-state NMR;33–37 X-ray powder diffraction;38,39
g, with the same limitations and assumptions coming
thermally stimulated depolarization current;40,41 etc.
along. As discussed earlier, the experimental ver-
These techniques are useful in the study of solid
ification of the solubility values resulting from these
dispersions at molecular level and the mechanistic
modeling studies is difficult, especially at tempera-
investigation of the drug–polymer interactions. On
tures close to and below Tg, and some inconsistency
the other hand, these techniques focus on the static
was also observed between different studies.29
mixing state of the system (i.e., whether the two
In conclusion, the current research on the deter-
components are homogeneously mixed) and they do
mination of drug–polymer miscibility and solubility is
not elucidate the thermodynamic nature of the
overall preliminary and the estimation at low
system (i.e., equilibrium miscibility).
temperatures is both experimentally and theoreti-
Similar to measuring drug solubility in polymers,
cally challenging. More fundamental understanding
the experimental determination of miscibility at
may be necessary for further progress.
temperatures below Tg is practically hard to achieve.
In addition, amorphous drugs usually tend to crystal-
lize, making the miscibility measurement even more PRACTICAL CONSIDERATIONS IN INDUSTRIAL
difficult. Experimental data obtained at tempera- DEVELOPMENT
tures above Tg coupled with modeling have therefore
been shown useful in the study of the thermody- Currently, the development of solid dispersions in the
namics of solid dispersions. Recent work done by industry remains largely trial and error, where a
Marsac et al.25 applied the Flory–Huggins theory to variety of solid dispersions with different polymer

DOI 10.1002/jps JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 99, NO. 7, JULY 2010
2946 QIAN, HUANG AND HUSSAIN

carriers and drug loadings are prepared and eval- dispersions: (1) obtain Tg of the system and give the
uated in parallel, with physical stability being on top best estimation (e.g., extrapolation) of the drug
of the watching list. This approach, although feasible solubility in polymer across the relevant temperature
in theory, is both time and resource-consuming. A range, (2) keep the drug loading in solid dispersion
rational approach that enables the selection of solid below the solubility at Tg (i.e., C1 in Fig. 1) as long as
dispersion formulation and storage conditions would the dose requirement can be satisfied, (3) in cases of
help streamlining the industrial practice. high-dose requirement, give the best estimation of
Figure 1 might be useful to guide a rational drug miscibility in polymer and use miscibility at Tg
approach for comprehensive comparison of different (C2 in Fig. 1) as the upper limit of drug loading,
solid dispersions side by side based on their physical (4) evaluate the hygroscopicity of the solid dispersion
stability. To use this figure, the major difficulties to address the impact of relevant water content,
could be the correct estimation of the two thermo- (5) use information from 1 to 4 combined with other
dynamic parameters, that is, drug–polymer misci- considerations (e.g., dissolution behavior, in vivo
bility and solubility, as well as the possible performance, manufacturing feasibility, etc.) to select
complications with regard to the correlation between appropriate polymer; (6) manufacture and store solid
these parameters and the physical stability of a solid dispersion within thermodynamically or kinetically
dispersion. On the other hand, kinetics plays a heavy stable zones (i.e., Zones I, II, III, and V in Fig. 1)
role in the practical development of solid dispersions. whenever possible, and (7) study the crystallization
Typical engineering processes, like spray drying and kinetics of the drug and the mixture (out of the scope
hot-melt extrusion, are capable of producing well of this commentary).
mixed systems, which remain as homogeneous With the complexity and controversy around the
dispersions (e.g., single Tg from DSC) kinetically determination and implication of drug–polymer
stable at temperature below its Tg. In a spray drying solubility and miscibility, more fundamental investi-
process, drug and polymer are dissolved in a common gations are highly desirable to reveal the details of the
solvent and rapidly dried (1 s). As long as appro- destabilization mechanism of amorphous solid dis-
priate solvent and processing conditions are applied, persions and then some industrial standards may be
initial spray dried dispersions would mostly be well established for solid dispersion development.
mixed even if the miscibility limit is exceeded.
Similarly in a well-designed hot-melt extrusion
process, drug is dissolved in polymer at high ACKNOWLEDGMENTS
temperatures (typically close to the melting of the
drug), mixed well by screws and the resultant viscous The authors thank Prof. Lian Yu and Dr. Jing Tao
solution is quenched to room temperature to form a (University of Wisconsin, Madison) for their valuable
homogeneous mixture. These systems, although inputs and helpful discussions.
might be thermodynamically unstable, may still be
proved kinetically stable enough as long as the mixing
status between drug and polymer can be confirmed REFERENCES
and maintained after manufacturing. The kinetic
effect on the stability of solid dispersions is out of the 1. Hancock BC, Zografi G. 1997. Characteristics and significance
of the amorphous state in pharmaceutical systems. J Pharm Sci
scope of this commentary. However, it certainly has 86:1–12.
very critical impact on the physical stability and 2. Yu L. 2001. Amorphous pharmaceutical solids: Preparation,
requires in-depth exploration. characterization and stabilization. Adv Drug Deliv Rev 48:27–
Using Figure 1 in practical applications, one has to 42.
keep in mind that moisture is ubiquitous and water is 3. Hancock BC, Shamblin SL, Zografi G. 1995. Molecular mobility
of amorphous pharmaceutical solids below their glass transi-
always the third component in any solid dispersion. It tion temperatures. Pharm Res 12:799–806.
is well known that water can bring profound impact 4. Hancock BC. 2002. Disordered drug delivery: Destiny,
on Tg, solubility, and miscibility. The Tg of a ‘‘wet’’ dynamics and the Deborah number. J Pharm Pharmacol 54:
solid dispersion can be obtained experimentally with 737–746.
some careful experimental control,28,30,31 or be 5. Sekiguchi K, Obi N, Ueda Y. 1964. Studies on absorption of
eutectic mixture. I. A comparison of the behavior of eutectic
estimated using the Fox equation, with the knowl- mixture of sulfathiazole and that of ordinary sulfathiazole in
edge of water content and Tg of pure water (108 to man. Chem Pharm Bull (Tokyo) 9:866–872.
1378C).44 Depending on the hygroscopicity of a 6. Sekiguchi K, Obi N, Ueda Y. 1964. Studies on absorption of
drug–polymer system, the moisture content may be eutectic mixture. II. Absorption of fused conglomerates of
critical for the drug loading limit and stability of the chloramphenicol and urea in rabbits. Chem Pharm Bull (Tokyo)
12:134–144.
formulation. 7. Repka MA, Majumdar S, Kumar BS, Srirangam R, Upadhye
In conclusion, a few practical notifications can be SB. 2008. Applications of hot-melt extrusion for drug delivery.
adopted for the evaluation and development of solid Expert Opin Drug Deliv 5:1357–1376.

JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 99, NO. 7, JULY 2010 DOI 10.1002/jps
DRUG–POLYMER SOLUBILITY AND MISCIBILITY 2947

8. Leuner C, Dressman J. 2000. Improving drug solubility for oral cular level solid dispersions containing poly(vinylpyrrolidone)
delivery using solid dispersions. Eur J Pharm Biopharm 50:47– and sorbed water. Pharm Res 25:647–656.
60. 29. Tao J, Sun Y, Zhang GG, Yu L. 2009. Solubility of small-
9. Fatouros DG, Karpf DM, Nielsen FS, Mullertz A. 2007. Clinical molecule crystals in polymers: D-mannitol in PVP, indometha-
studies with oral lipid based formulations of poorly soluble cin in PVP/VA, and nifedipine in PVP/VA. Pharm Res 26:855–
compounds. Ther Clin Risk Manag 3:591–604. 864.
10. Gao P, Morozowich W. 2006. Development of supersaturatable 30. Vasanthavada M, Tong WQ, Joshi Y, Kislalioglu MS. 2004.
self-emulsifying drug delivery system formulations for improv- Phase behavior of amorphous molecular dispersions. I: Deter-
ing the oral absorption of poorly soluble drugs. Expert Opin mination of the degree and mechanism of solid solubility.
Drug Deliv 3:97–110. Pharm Res 21:1598–1606.
11. Chen Y, Zhang GG, Neilly J, Marsh K, Mawhinney D, Sanzgiri 31. Vasanthavada M, Tong WQ, Joshi Y, Kislalioglu MS. 2005.
YD. 2004. Enhancing the bioavailability of ABT-963 using solid Phase behavior of amorphous molecular dispersions. II: Role of
dispersion containing Pluronic F-68. Int J Pharm 286:69–80. hydrogen bonding in solid solubility and phase separation
12. Law D, Krill SL, Schmitt EA, Fort JJ, Qiu Y, Wang W, Porter kinetics. Pharm Res 22:440–448.
WR. 2001. Physicochemical considerations in the preparation 32. Krause S, Iskander M. 1977. Phase separation in styrene-a-
of amorphous ritonavir-poly(ethylene glycol) 8000 solid disper- methyl styrene block copolymers. In: Klempner D, Frisch K,
sions. J Pharm Sci 90:1015–1025. editors. Polymer science and technology. New York: Plenum
13. Zerrouk N, Chemtob C, Arnaud P, Toscani S, Dugue J. 2001. Pub Corp 10th edition. pp 231–243.
In vitro and in vivo evaluation of carbamazepine-PEG 6000 33. Aso Y, Yoshioka S, Miyazaki T, Kawanishi T, Tanaka K,
solid dispersions. Int J Pharm 225:49–62. Kitamura S, Takakura A, Hayashi T, Muranushi N. 2007.
14. Qian F, Tao J, Desikan S, Hussain M, Smith RL. 2007. Mechan- Miscibility of nifedipine and hydrophilic polymers as measured
istic investigation of Pluronic based nano-crystalline drug– by (1)H-NMR spin-lattice relaxation. Chem Pharm Bull
polymer solid dispersions. Pharm Res 24:1551–1560. (Tokyo) 55:1227–1231.
15. Angell CA. 1995. Formation of glasses from liquids and biopo- 34. Berendt RT, Sperger DM, Munson EJ, Isbester PK. 2006. Solid-
lymers. Science 267:1924–1935. state NMR spectroscopy in pharmaceutical research and ana-
16. Rumondor AC, Stanford LA, Taylor LS. 2009. Effects of poly- lysis. TrAC Trends Anal Chem 25:977–984.
mer type and storage relative humidity on the kinetics of 35. Nelson BN, Schieber LJ, Barich DH, Lubach JW, Offerdahl TJ,
felodipine crystallization from amorphous solid dispersions. Lewis DH, Heinrich JP, Munson EJ. 2006. Multiple-sample
Pharm Res 26:2599–2606. probe for solid-state NMR studies of pharmaceuticals. Solid
17. Rumondor AC, Ivanisevic I, Bates S, Alonzo DE, Taylor LS. State Nucl Magn Reson 29:204–213.
2009. Evaluation of drug–polymer miscibility in amorphous 36. Lubach JW, Padden BE, Winslow SL, Salsbury JS, Masters
solid dispersion systems. Pharm Res 26:2523–2534. DB, Topp EM, Munson EJ. 2004. Solid-state NMR studies of
18. Taylor LS, Zografi G. 1997. Spectroscopic characterization of pharmaceutical solids in polymer matrices. Anal Bioanal Chem
interactions between PVP and indomethacin in amorphous 378:1504–1510.
molecular dispersions. Pharm Res 14:1691–1698. 37. Geppi M, Guccione S, Mollica G, Pignatello R, Veracini CA.
19. Campbell AN, Campbell AJR. 1937. Concentrations, total and 2005. Molecular properties of ibuprofen and its solid disper-
partial vapor pressures, surface tensions and viscosities, in the sions with Eudragit RL100 studied by solid-state nuclear mag-
systems phenol–water and phenol–water–4% succinic acid. netic resonance. Pharm Res 22:1544–1555.
J Am Chem Soc 59:2481–2488. 38. Newman A, Engers D, Bates S, Ivanisevic I, Kelly RC,
20. Rubinstein M, Colby RH. 2003. Thermodynamics of blends and Zografi G. 2008. Characterization of amorphous API:polymer
solutions. In: Polymer physics. New York: Oxford University mixtures using X-ray powder diffraction. J Pharm Sci 97:4840–
Press Inc,. pp 137–170. 4856.
21. Bernstein J, Davey RJ, Henck JO. 1999. Concomitant poly- 39. Ivanisevic I, Bates S, Chen P. 2009. Novel methods for the
morphs. Angew Chem Int Ed Engl 38:3440–3461. assessment of miscibility of amorphous drug–polymer disper-
22. Flory PJ. 1941. Thermodynamics of high polymer solutions. sions. J Pharm Sci 98:3373–3386.
J Chem Phys 9:660–661. 40. Shmeis RA, Wang Z, Krill SL. 2004. A mechanistic investiga-
23. Huggins ML. 1942. Thermodynamic properties of solutions of tion of an amorphous pharmaceutical and its solid dispersions.
long-chain compounds. Ann N Y Acad Sci 43:1–32. Part I: A comparative analysis by thermally stimulated depo-
24. Crowley KJ, Zografi G. 2002. Water vapor absorption into larization current and differential scanning calorimetry.
amorphous hydrophobic drug/poly(vinylpyrrolidone) disper- Pharm Res 21:2025–2030.
sions. J Pharm Sci 91:2150–2165. 41. Shmeis RA, Wang Z, Krill SL. 2004. A mechanistic investiga-
25. Marsac PJ, Shamblin SL, Taylor LS. 2006. Theoretical and tion of an amorphous pharmaceutical and its solid dispersions.
practical approaches for prediction of drug–polymer miscibility Part II: Molecular mobility and activation thermodynamic
and solubility. Pharm Res 23:2417–2426. parameters. Pharm Res 21:2031–2039.
26. Marsac PJ, Li T, Taylor LS. 2009. Estimation of drug–polymer 42. Koningsveld R. 1970. Liquid–liquid equilibria in multi-
miscibility and solubility in amorphous solid dispersions using component polymer systems. Discuss Faraday Soc 49:144–
experimentally determined interaction parameters. Pharm Res 161.
26:139–151. 43. Koningsveld R, Kleintjens LA, Shultz AR. 1970. Liquid–liquid
27. Gordon M, Taylor J. 1952. Ideal copolymers and the second- phase separation in multicomponent polymer solutions—9.
order transitions of synthetic rubbers. I. Noncrystalline copo- J Polym Sci Part A 2 Polym Phys 8:1261–1278.
lymers. J Appl Chem 2:493–500. 44. Velikov V, Borick S, Angell C. 2001. The glass transition
28. Marsac PJ, Konno H, Rumondor AC, Taylor LS. 2008. Recrys- of water, based on hyperquenching. Science 294:2335–
tallization of nifedipine and felodipine from amorphous mole- 2338.

DOI 10.1002/jps JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 99, NO. 7, JULY 2010

You might also like