Vous êtes sur la page 1sur 10

ADR-12895; No of Pages 10

Advanced Drug Delivery Reviews xxx (2016) xxx–xxx

Contents lists available at ScienceDirect

Advanced Drug Delivery Reviews

journal homepage: www.elsevier.com/locate/addr

Recent advances in co-amorphous drug formulations☆


Swapnil Jayant Dengale a, Holger Grohganz b, Thomas Rades b,⁎, Korbinian Löbmann b
a
Manipal College of Pharmaceutical Sciences, Manipal University, Manipal, Karnataka, India
b
Department of Pharmacy, University of Copenhagen, Copenhagen, Denmark

a r t i c l e i n f o a b s t r a c t

Article history: Co-amorphous drug delivery systems have recently gained considerable interest in the pharmaceutical field be-
Received 6 October 2015 cause of their potential to improve oral bioavailability of poorly water-soluble drugs through drug dissolution en-
Accepted 9 December 2015 hancement as a result of the amorphous nature of the material. A co-amorphous system is characterized by the
Available online xxxx
use of only low molecular weight components that are mixed into a homogeneous single-phase co-amorphous
blend. The use of only low molecular weight co-formers makes this approach very attractive, as the amount of
Keywords:
Co-amorphous
amorphous stabilizer can be significantly reduced compared with other amorphous stabilization techniques. Be-
Solid dispersion cause of this, several research groups started to investigate the co-amorphous formulation approach, resulting in
Poorly soluble drugs an increasing amount of scientific publications over the last few years. This study provides an overview of the co-
Molecular interactions amorphous field and its recent findings. In particular, we investigate co-amorphous formulations from the view-
Increased dissolution point of solid dispersions, describe their formation and mechanism of stabilization, study their impact on disso-
lution and in vivo performance and briefly outline the future potentials.
© 2016 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2. What are co-amorphous formulations? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3. Technologies for the preparation of co-amorphous systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4. Mechanism of physical stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.1. Amorphous solubility. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.2. Glass transition temperature (Tg) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.3. Intermolecular interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.4. Intimate mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5. Dissolution properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6. In vivo performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
7. Concluding remarks and future outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0

1. Introduction their crystalline counterpart [1]. The main drawback of using pure
amorphous highly soluble drugs is their physical instability with respect
The development of amorphous drug delivery systems has been to their inherent tendency to recrystallize into the poorly soluble crys-
widely investigated in academia and by the pharmaceutical industry talline form as they are thermodynamically unstable [2].
to overcome the poor aqueous solubility of many drugs. Briefly, the As only pure amorphous drugs often appear non-feasible in drug de-
same solid material can be crystalline or amorphous, where amorphous livery systems, a major focus within amorphous research and develop-
drugs exhibit a significantly higher solubility and dissolution rate than ment is the stability of the amorphous form through the use of
excipients. Several approaches have been introduced in previous stud-
ies, including polymer-based glass solutions, mesoporous silica and
☆ This review is part of the Advanced Drug Delivery Reviews theme issue on “Amorphous
pharmaceutical solids”.
co-amorphous formulations. Of them, the co-amorphous strategy has
⁎ Corresponding author. recently gained considerable interest in the pharmaceutical field as it
E-mail address: thomas.rades@sund.ku.dk (T. Rades). provides opportunities to overcome shortcomings associated with the

http://dx.doi.org/10.1016/j.addr.2015.12.009
0169-409X/© 2016 Elsevier B.V. All rights reserved.

Please cite this article as: S.J. Dengale, et al., Recent advances in co-amorphous drug formulations, Adv. Drug Deliv. Rev. (2016), http://dx.doi.org/
10.1016/j.addr.2015.12.009
2 S.J. Dengale et al. / Advanced Drug Delivery Reviews xxx (2016) xxx–xxx

other two approaches. The objective of this study is to provide an over-


view on the state of the art of co-amorphous drug formulations.

2. What are co-amorphous formulations?

Polymer-based glass solutions, mesoporous silica and co-amorphous


formulations all come under the term glass solutions, which itself is a Scheme 1. Classification of co-amorphous mixtures within glass solutions based on the
subcategory of solid dispersions. The use of this expression is very in- choice of the stabilizing agent.

consistent in the pharmaceutical field; therefore, this section provides


a short guidance to the complex classification of solid dispersions. In mesoporous silica-based glass solutions, the drugs are
The term solid dispersion was defined by Chiou and Riegelman [3] in amorphized by adsorption onto the surface of the silica particles,
1971 as “a dispersion of one or more active ingredients in an inert carri- which consists of a matrix of pores with diameter between 2 and
er at the solid state prepared by the melting (fusion), solvent, or 50 nm [15]. On the one hand, stabilization of the amorphous drug is
melting-solvent method.” According to this definition, solid dispersions achieved through molecular interactions between the drug and the
can be classified according to their number of solid-state phases and the functional groups of the silica matrix [16,17]. On the other hand, crystal-
physical state of these phases. As presented in Table 1, solid dispersions lization is inhibited physically by the pore diameter of the materials,
can be very diverse, including eutectic mixtures, solid solutions, glass which may be smaller than the size of a crystal nucleus of the drug
solutions and glass suspensions. A more detailed description of the dif- [18]. However, the main drawbacks of mesoporous silica-based glass so-
ferent types of solid dispersions is provided by Chiou and Riegelman [3] lutions are their production, which predominantly involves the use of
and Laitinen et al. [4] organic solvents for drug loading [18], and often a limited loading capac-
Glass solutions, that is, single amorphous phase systems, are often ity of only 20%–30% [19].
referred to as amorphous solid dispersions (ASD) and can further be The co-amorphous drug formulation approach is characterized by
subdivided according to the excipients that stabilize the amorphous the combination of two or more low molecular weight components
drug. Vaka et al. differentiated them into polymeric and non- that form a homogeneous amorphous single-phase system [20,21]. In
polymeric excipients, where the former can further be classified into order to differentiate glass solutions comprising only small molecules
ionic and non-ionic polymers [5]. The group of non-polymeric excipi- from those with stabilizing polymers or mesoporous silica matrices,
ents can further be divided into mesoporous silica-based glass solutions Chieng et al. coined the term ‘co-amorphous’ in 2009 [22]. Using this ap-
and those containing only low molecular weight components, the so- proach, it was proposed that the amount of stabilizing excipient (if
called co-amorphous formulations (Scheme 1). used) can be drastically reduced due to the low molecular weight of
To date, polymer-based ASDs are by far the most investigated ASDs. the co-amorphous co-former. Two types of co-amorphous principles
Thus, the term solid dispersion is highly connected to the use of poly- have been proposed so far, namely drug–drug combinations and
mers in stabilizing amorphous drugs. This is misleading as solid disper- drug–excipient mixtures [21]. In the first type, two pharmacologically
sions are a rather large group of different types of solid mixtures relevant drugs intended for multidrug therapies are combined, where
(Table 1). Thus, for a clearer differentiation, ASDs with polymeric excip- both drugs stabilize each other in the amorphous form. Thus, both
ients are called polymer-based glass solutions in this study. drugs act as an active component and stabilizing excipient at the same
In polymer-based glass solutions, polymeric carriers are used to sta- time. As a result of the stable amorphous system, both of the poorly sol-
bilize the amorphous drug and improve its solubility and dissolution uble drugs achieve a higher solubility and dissolution rate. In the second
rate [6]. In these systems, stabilization of the drug in its amorphous type, low molecular weight excipients such as amino acids are used to
form is achieved by several factors. One key aspect is the solubility of prepare stable and quick-dissolving co-amorphous drug–excipient
the drug in the amorphous polymer [7,8]. Below its limit of solubility, blends.
the drug is molecularly dispersed in the amorphous polymer and stabi-
lized by physical separation of the molecules between the polymer 3. Technologies for the preparation of co-amorphous systems
chains. Most polymeric carriers have a high glass transition temperature
(Tg), and thus increase the Tg (while reducing the molecular mobility) Several techniques for the production of amorphous drugs have
of the drug in the glass solution compared to that in its pure amorphous been described in the literature. Depending on the mechanism involved,
form [9]. Furthermore, intermolecular interactions between the drug these techniques can be divided into two types as including thermody-
and functional groups of the polymer have been found to play a role namic and kinetic disordering processes [23,24]. The thermodynamic
in the stabilization mechanism [10]. However, limited drug solubility pathway has a thermodynamically stable non-crystalline form as a
in large polymeric excipients often increases the dosage and does not starting point, that is, the drug as a melt or in solution. In order to obtain
necessarily make the formulation very stable against recrystallization the amorphous drug, the melt needs to be subsequently vitrified by
[11–13]. Another drawback is the hygroscopic nature of many polymer- rapid cooling, a process called quenching, or the drug needs to be pre-
ic carriers, which results in absorption of moisture. The absorbed mois- cipitated from the solution, followed by solvent removal. The kinetic
ture acts as a plasticizer, thus reducing the Tg and increasing mobility, pathway involves direct solid-state conversion of the crystalline drug
which in turn can result in phase separation and recrystallization [13]. into its amorphous form. This can be achieved by continuously intro-
Thus, despite an active research interest, polymer-based glass solutions ducing crystal defects and disorders through shear forces, crushing
have led to only a few marketed products [5,14]. and impact during a milling process [2,25]. In general, both mechanisms
have been described for the preparation of amorphous blends [26].
The co-amorphous drug formulation approach is still in its early
Table 1
stage of development, hence majority of studies focused on the basic
Classification of solid dispersions (reproduced from ref. [4]).
understanding of these systems using laboratory-scale preparative
Solid dispersion Number of phases Physical state of phase(s) techniques such as quenching [11,27–37], solvent evaporation [38–47]
Eutectic mixture 2 C/C and ball milling [28,43,47–57]. All of these techniques are attractive as
Solid solution 1 C they represent fast and easy ways of (co-)amorphization, and are ideal
Glass solution 1 A for screening purposes as only small sample sizes are required. In addi-
Glass suspension 2 A/A or A/C
tion, quenching offers the possibility to quickly assess the critical

Please cite this article as: S.J. Dengale, et al., Recent advances in co-amorphous drug formulations, Adv. Drug Deliv. Rev. (2016), http://dx.doi.org/
10.1016/j.addr.2015.12.009
S.J. Dengale et al. / Advanced Drug Delivery Reviews xxx (2016) xxx–xxx 3

physicochemical parameters such as Tg, miscibility and recrystalliza- those in the amorphous glassy state. However, although molecules in
tion, as the co-amorphous systems can be prepared and analyzed direct- the glassy state are kinetically frozen, they still exhibit motion, albeit
ly in situ within a differential scanning calorimeter (DSC). However, all at a much lower rate. This phenomenon is called relaxation, which
these methods cannot be applied for any type of drug or co-forming ex- helps an amorphous material to crystallize over time into a thermody-
cipient. Quenching is applicable only to compounds that do not degrade namically stable form even at temperatures much below its Tg. In
upon melting. Solvent evaporation can be challenging if the poorly order to keep a glassy material in its amorphous form, it is suggested
water-soluble components also show poor solubility in organic sol- to store the material at least 50° below the Tg [2,9,64,66,69].
vents, or if the two components do not dissolve in the same solvent in In glass solutions, the Tg of the amorphous multi-component system
appropriate concentrations, that is, only one component dissolves in or- is usually found between the Tgs of the individual components. This re-
ganic solvents and the co-amorphous co-former dissolves in aqueous lationship is described by the Gordon–Taylor equation as follows:
solvents. On the contrary, ball milling might not be satisfactorily effi-
cient for the disruption of crystal lattice and thus, might not result in a w1 T g1 þ Kw2 T g2
T g12 ¼
complete (co-)amorphization. Therefore, the physicochemical proper- w1 þ Kw2
ties of drugs and excipients usually determine the preparative
technique. where Tg12 is the Tg of the amorphous mixture, Tg1 and Tg2 represent the
With a view of developing industrially more feasible production Tgs of the respective individual components, w1 and w2 represent the
methods, a few studies have reported the use of scalable techniques respective weight fractions and K is a constant. It is thus not surprising
for the preparation of co-amorphous formulations including spray- that the inclusion of drugs into polymeric carriers with high Tg, such
drying [58], freeze-drying [59] and ultrasound extrusion [60,61]. Fur- as PVP, has been shown to improve physical stability due to the in-
thermore, inkjet printing has been used for the preparation of co- creased Tg of the polymer–drug blend compared with the Tg of the
amorphous indomethacin–arginine systems to obtain fabricated pure amorphous drug [25]. This anti-plasticizing effect of polymers is
printed systems that allow flexible and more individualized dosing one of the key characteristics of polymer-based glass solutions.
and thus, the development of quick-dissolving personalized medicines Compared with polymeric excipients, low molecular weight compo-
[62]. nents, that is, majority of drugs, usually have a comparatively low Tg. As
co-amorphous systems contain only low molecular weight molecules,
4. Mechanism of physical stabilization the possibility of anti-plasticization is limited. Nevertheless, this princi-
ple has been shown in several cases of co-amorphous formulations. In
4.1. Amorphous solubility particular, the use of amino acids as co-amorphous excipients has
been shown to result in relatively high Tgs in co-amorphous blends,
As mentioned in Section 2, a co-amorphous system is a single-phase for example, with the drugs carbamazepine and indomethacin [51].
amorphous mixture of two or more low molecular weight components. The developed co-amorphous binary and ternary drug–amino acid sys-
For a system to be able to from a single phase, the components in the tems exhibited excellent physical stability over at least 6 months,
blend need to exhibit complete miscibility in the amorphous form. For whereas the pure amorphous drugs recrystallized within 7 days. Not
thermostable compounds, the miscibility of the components in the mol- only the increase of Tg, but also the increased stability of the systems
ten state can easily be determined by the phase diagram [11,28,33,36]. are attributed to molecular interactions between the drugs and amino
By quenching these single-phase melts, one can obtain a single-phase acids [53]. In particular, tryptophan exhibited excellent co-forming
co-amorphous mixture. Similarly, Marsac et al. showed solubility and anti-plasticizing properties similarly to a co-amorphous system, be-
using a melting point depression approach, where miscible systems cause of its high Tg of approximately 140 °C [52,56]. Furthermore, for
show a depression of the melting point of the drug and immiscible or strong ionic interactions between the components, the Tg of the co-
partially miscible systems show little or no such depression [63]. Solu- amorphous systems can be much higher than those of the individual
bility parameters can be used to check miscibility of thermolabile com- components [53,70]. Hence, elevated Tg of co-amorphous mixtures
pounds in the amorphous blend [36,50,51]. Another indicator for the over their individual amorphous compounds has been discussed as
formation of a homogeneous single-phase co-amorphous blend, one of the factors for improved physical stability of these systems.
where both components are dissolved in each other, is the observation
of a single Tg [3,64–66]. By contrast, immiscible or partially miscible 4.3. Intermolecular interactions
components result in two-phase amorphous mixtures, thereby produc-
ing two Tgs [67]. Many studies showed that co-amorphous systems exhibit a higher
Correspondingly, for polymer-based glass solutions, the thermody- physical stability than the individual amorphous drugs. Because the Tg
namic solubility of the drug in the amorphous polymer has been de- of co-amorphous systems is usually found in between the Tgs of the in-
scribed as one of the primary reason for stability; however, many dividual components, Tg alone cannot explain the increase of physical
drugs possess only a limited solubility of the drug in the polymeric car- stability. However, such an increase is attributed to several factors,
rier (often ≤20 wt.%) [68]. When the drug is supersaturated in the poly- including molecular interactions between the components in the co-
mer, phase separation into drug-rich and/or polymer/excipient-rich amorphous mixture. The majority of studies on the co-amorphous for-
domains may occur, followed by a rapid nucleation and crystal growth. mulation approach have attributed the physical stability of such sys-
Similarly, partially miscible or immiscible co-amorphous mixtures that tems to intermolecular interactions such as hydrogen bonding and/or
form a homogeneous phase initially after preparation might show fast π–π interactions [17,22,28,29,36,38,39,41–43,46,51,57].
phase separation and crystallization. In individual amorphous components, the molecules are often ar-
ranged in a certain short-range molecular order, which is reflected in
4.2. Glass transition temperature (Tg) molecular interactions between like molecules, such as the formation
of homodimers in amorphous indomethacin or naproxen [37]. The
Tg of an amorphous material is defined as the temperature at which homodimers are often found similarly in the crystalline state of the
the material transforms from its glassy state to a supercooled liquid drugs and thus, recrystallization in the pure amorphous form occurs
state upon heating [2,25]. At the Tg, the material changes from a solid- usually at a rather high rate. On the contrary, in the co-amorphous
like form to viscous liquid-like form, thereby drastically changing its blend, this molecular short-range order gets disturbed in favor of the
molecular mobility. Because of the higher molecular mobility, materials formation of intermolecular interactions between dissimilar molecules,
in the supercooled liquid state crystallize at a much higher rate than that is, the two different components in the co-amorphous blend. Such a

Please cite this article as: S.J. Dengale, et al., Recent advances in co-amorphous drug formulations, Adv. Drug Deliv. Rev. (2016), http://dx.doi.org/
10.1016/j.addr.2015.12.009
4 S.J. Dengale et al. / Advanced Drug Delivery Reviews xxx (2016) xxx–xxx

formation of heterodimers is shown in several studies of co-amorphous consider using the interacting adduct between the components in
formulations, such as naproxen/indomethacin, ritonavir/quercetin and amorphous blends rather than the pure components when using the
cimetidine/piroxicam systems [36,42,46]. For such a system to recrys- Gordon–Taylor equation. On the contrary, when finding the largest de-
tallize, it is suggested that the intermolecular bonds within the hetero- viation from the classical Gordon–Taylor approach by determining the
dimer must be broken, followed by a rearrangement of molecules to experimental Tgs over a range of different ratios, one might be able to
form homodimers and the subsequent establishment of long-range identify the co-amorphous blend with the largest degree of interactions
order in the crystal lattice. This cascade is ongoing for a considerably between the components in the mixture.
longer timescale, leading to a reduced likelihood of recrystallization
and thus, prolonged physical stability of the co-amorphous systems 4.4. Intimate mixing
[20,21,56]. Furthermore, the formation of even stronger ionic interac-
tions has been described for co-amorphous systems. Yamamura et al. In some co-amorphous systems, an improved physical stability was
prepared co-amorphous systems of cimetidine with indomethacin and observed as a result of intimate mixing, that is, physical separation of
diflunisal, and found salt formation between the imidazole ring of ci- similar molecules in the homogeneous co-amorphous blend. Simvastat-
metidine and carboxyl groups of indomethacin and diflunisal [38,39]. in/glipizide systems showed improved physical stability over the indi-
Strong molecular interactions have also been shown to occur in vidual amorphous drugs without any detectable intermolecular
drug–excipient mixtures [11,30,32,47]. In particular, the use of amino interactions or an increased Tg [50]. Dengale et al. reported similar ob-
acids as excipients has recently shown strong potential to stabilize servations for ritonavir/indomethacin systems after co-precipitation
drugs in the co-amorphous form [51–53,56,58,70]. This approach was followed by solvent evaporation [45]. The recrystallization of these sys-
originally based on the assumption that drugs interact at the molecular tems is suspected to be a result of slow demixing and phase separation.
level with amino acids at their respective target sites (receptor proteins) In general, for most co-amorphous systems, a clear separation of the
in the body and thus, may also be able to interact with amino acids in a aforementioned stabilization mechanisms is not possible, but the in-
co-amorphous mixture [51]. For this purpose, binary and ternary co- creased stability is rather the result of a combination of these mecha-
amorphous systems comprising indomethacin and carbamazepine nisms. Together with molecular interactions, Löbmann et al., for
with a set of amino acids (receptor and non-receptor) were prepared example, attributed the physical stability of co-amorphous systems of
by ball milling and found to be stable for at least 6 months at 40 °C. indomethacin and carbamazepine with various amino acids to the
The physical stability of the indomethacin/arginine system was attribut- molecular-level mixing of drug with amino acids, their molecular inter-
ed to ionic interactions between the carboxylic acid group of indometh- actions and the increased Tg after ball milling [51,53]. It is important to
acin and the guanidine moiety of arginine. For the tryptophan/ mention that solid-state solubility of the components is the first re-
carbamazepine system, hydrogen bonding and π–π interactions have quirement for a co-amorphous blend, whereas increased Tg, molecular
been found responsible for the increased physical stability. However, interactions and intimate mixing occur due to the miscibility of the
obvious differences exist between the anticipated interactions based components in the amorphous blend. Table 2 summarizes the factors
on in vivo binding and actual interactions in the co-amorphous mix- responsible for the increased stability of different co-amorphous
tures. Unlike in vivo binding where only the side chains of the amino formulations.
acids are able to interact with the drug, it was found that the whole
amino acid molecule, that is, side chain as well as head group interacted 5. Dissolution properties
with the drug in the co-amorphous states [21,51].
Such intermolecular interactions in the solid state have been inves- Because of their higher internal energy, amorphous phases possess a
tigated by Fourier transform infrared spectroscopy [28,30,36,40,41,43, higher solubility and dissolution rate than their crystalline counterparts
45–47,49,52–54,71] and solid-state nuclear magnetic resonance [2,20,65,76]. Accordingly, co-amorphous systems have been found to
[32–34,38,39,42,56]. Furthermore, a deviation of the experimental Tgs show improved dissolution behavior not only over their crystalline
from theoretical Tgs following the Gordon–Taylor equation can be counterparts, but moreover also over their individual amorphous
used for identifying molecular interactions. The Gordon–Taylor equa- forms (Table 3) [36,45,47,50–52,55,57]. For example, Allesø et al. ob-
tion describes the Tg of a homogeneous amorphous blend of two com- served a higher dissolution rate of co-amorphous naproxen/cimetidine
ponents and is based on two important assumptions: (1) ideal free than both crystalline drugs and amorphous cimetidine (a comparison
volume additivity of the two components in the amorphous mixture to amorphous naproxen was not made because of its high instability
and (2) no specific interactions exist between these components and fast recrystallization immediately after preparation). The dissolu-
(ideal mixing behavior). As the Gordon–Taylor equation explains the tion rate of pure amorphous cimetidine was found to be identical to
dependence of Tg on the composition of the amorphous blends with that of crystalline cimetidine, indicating its recrystallization upon con-
the aforementioned assumptions, deviations of the calculated values tact with the dissolution medium. However, when co-milled with
from experimentally observed ones have been interpreted as the possi- naproxen, cimetidine showed a 2-fold increase in dissolution rate with-
bility of intermolecular interactions [64,72–75]. out any evidence of recrystallization. The authors suggested that co-
A modified use of the Gordon–Taylor equation has been described amorphization prevented cimetidine from recrystallization upon
for co-amorphous naproxen/indomethacin when the degree of molecu- dissolution [57]. Similarly, fast solvent-mediated recrystallization was
lar interactions, for example, the formation of a heterodimer, between observed for pure amorphous lurasidone HCl during dissolution,
the components is known [36]. In such a case, the deviation of the ex- which resulted in a faster offset of the dissolution rate than the dissolu-
perimentally determined Tgs from the theoretical Tgs calculated using tion rate of crystalline lurasidone HCl (Fig. 1). Recrystallization was con-
the Gordon–Taylor equation was largest for the co-amorphous system firmed as birefringence on the surface of the intrinsic dissolution
showing the largest degree of molecular interactions, that is, the hetero- compact. On the contrary, co-amorphous lurasidone HCl/saccharine
dimer at 1:1 M ratio. Assuming that the heterodimer is made of one showed a continuous fast dissolution (5.6-fold faster than crystalline
component, whereas any excess drug represents the second component lurasidone HCl) for the whole duration of the dissolution experiment
in the co-amorphous blends, the theoretical Tgs of the mixtures can be [47]. Unlike pure amorphous lurasidone HCl, birefringence was not ob-
recalculated. In the above example, an ideal fit of the experimental served on the surface of the co-amorphous tablet, indicating the ab-
values with the theoretical ones using this modified use of the sence of recrystallization during dissolution. Thus, co-amorphization
Gordon–Taylor equation was obtained. This finding suggested that does not only increase the dissolution rate of the drug, but can also
ideal mixing behavior was observed for the heterodimer and any excess help prevent solvent-induced recrystallization upon dissolution. Again,
drug in those mixtures. Therefore, it has been suggested that one should molecular interactions played a crucial role in this stabilization

Please cite this article as: S.J. Dengale, et al., Recent advances in co-amorphous drug formulations, Adv. Drug Deliv. Rev. (2016), http://dx.doi.org/
10.1016/j.addr.2015.12.009
S.J. Dengale et al. / Advanced Drug Delivery Reviews xxx (2016) xxx–xxx 5

Table 2 proline mixtures. Jensen et al. applied two formulation principles.


Factors increasing physical stability of co-amorphous systems. First, the amorphous form of naproxen was stabilized by co-
System Main type of stabilization Reference amorphization with the high-Tg amino acid tryptophan (see
Cimetidine/Indomethacin Salt formation [39]
Section 4.2) or the strongly interacting amino acid arginine (salt forma-
Cimetidine/Diflunisal Salt formation [38] tion). Second, dissolution improvement was achieved by inclusion of
γ-Indomethacin/Ranitidine hydrochlorideHydrogen bonding [22] the highly soluble amino acid proline, which itself is a poor amorphous
Naproxen/Cimetidine π–π interaction [57] stabilizer [52]. As a result, both ternary mixtures exhibited an increased
Cimetidine/Piroxicam Hydrogen bonding [42]
dissolution rate than the binary naproxen/tryptophan and naproxen/ar-
γ-Indomethacin/Naproxen Hydrogen bonding [36]
Simvastatin/Glipizide Intimate mixing [50] ginine mixtures. However, in a subsequent study, it was shown that the
Ritonavir/Indomethacin Intimate mixing [45] highly soluble co-formers can facilitate the dissolution of a given drug
Acyclovir/Citric acid Hydrogen bonding [41] only to a certain degree [70]. When solubility of the co-former is too
Salt formation and Tg high and/or interactions between the components are not strong
Indomethacin/Arginine [51,53]
increase
Hydrogen bonding, π–π
enough, the highly soluble co-former from the co-amorphous system
Indomethacin/Tryptophan interaction and Tg [51,53] dissolves too fast, leaving the pure amorphous form in the suspension.
increase As a consequence, the drug loses its amorphous stabilizer and is thus,
Hydrogen bonding and prone to recrystallization.
Indomethacin/Phenylalanine [51,53]
π–π interaction
In order to investigate supersaturation of the drug from co-
Hydrogen bonding and
Indomethacin/Phenylalanine/Tryptophan
π–π interaction
[51,53] amorphous formulations, Heikinnen et al. studied the dissolution of dif-
Hydrogen bonding and ferent co-amorphous drug amino acid formulations under non-sink
Carbamazepine/Arginine/Tryptophan [51,53]
π–π interaction conditions in phosphate buffer and bio-relevent media, that is, fasted
Hydrogen bonding and (FaSSIF) and fed state-simulated intestinal fluid (FeSSIF) [55]. All of
Carbamazepine/Phenylalanine/Tryptophan [51,53]
π–π interaction
the investigated systems (simvastatin/lysine, glibenclamide/serine,
Hydrogen bonding and
Carbamazepine/Tryptophan [51,53] glibenclamide/threonine and glibenclamide/serine/threonine) showed
π–π interaction
Salt formation and an increased dissolution rate and long-lasting supersaturation com-
Indomethacin/Arginine/Phenylalanine [51,53]
hydrogen bonding pared with the individual amorphous and crystalline drugs (Fig. 2). In
Naproxen/Indomethacin Hydrogen bonding
the case of glibenclamide, the pure amorphous form provided a similar
Salt formation and
Naproxen/Arginine/Proline
hydrogen bonding
[52] supersaturation as the co-amorphous systems; however, the dissolution
Naproxen/Tryptophan Hydrogen bonding [52] rate of the pure amorphous form was slightly lower (Fig. 2(b)). Provid-
Naproxen/Tryptophan/Proline Hydrogen bonding [52] ed the pure amorphous form exhibited a performance similar to the co-
Naproxen/Arginine Salt formation [52] amorphous blends in this example, it is worth mentioning that the for-
Paracetamol/Citric acid Hydrogen bonding [32]
mation of a co-amorphous form was still beneficial because of the
Indomethacin/Tryptophan π–π interaction [56]
Furosemide/Tryptophan Hydrogen bonding [56] prolonged physical stability of amorphous glibenclamide [54]. Similarly,
Ritonavir/Quercetin Hydrogen bonding [46] Shayanfar et al. showed a long-lasting supersaturation of the two drugs
Lurasidone Hydrochloride/Saccharin Hydrogen bonding [47] atorvastatin and glibenclamide from their respective co-amorphous
Repaglinide/Saccharin Hydrogen bonding [43]
mixtures [40].
Atorvastatin/Carvedilol Hydrogen bonding [40]
Atorvastatin/Glibenclamide Hydrogen bonding [40]
Overall, the dissolution of co-amorphous formulations has been
Curcumin/Artemisin Hydrogen bonding [44] shown to be improved compared with the pure crystalline and amor-
phous drugs. Molecular interactions are again not only proved to be im-
portant for preventing solvent-induced recrystallization, but also
mechanism [36,57,70]. On the contrary, Dengale et al. found that ritona- contributed to a possible synchronized release of components from
vir/indomethacin systems showed an increase in the dissolution rate of the co-amorphous mixtures. The dissolution rate of the poorly soluble
ritonavir due to intimate mixing without intermolecular interactions drug from the co-amorphous blend is dependent on the solubility of
[45]. the co-former and the strength of the molecular interactions between
In addition to enhanced dissolution rates, some co-amorphous mix- drug and co-former. Furthermore, co-amorphous formulations can pro-
tures show a pairwise or synchronized dissolution of the individual vide long-lasting supersaturation.
components due to their short-range molecular order, that is, the for-
mation of heterodimers through hydrogen bonds [36,57]. For co- 6. In vivo performance
amorphous naproxen/cimetidine, it is suggested that pairwise solvation
and their interdependency upon dissolution were responsible for the In order to demonstrate the potential of the co-amorphous formula-
synchronized release. Both are a result of the intermolecular interac- tion approach, the positive in vitro dissolution results also need to be
tions at 1:1 M ratio. Similar findings were also reported for co- reflected in an increased in vivo bioavailability. Although the co-
amorphous naproxen/indomethacin, and again related to the intermo- amorphous technology is a rather new approach for improving amor-
lecular interactions between the two drugs in the mixture [36]. phous stability of drugs and their dissolution, only a limited number of
Because synchronized release of two components from the co- studies have been published on in vivo performance of these systems
amorphous blends appears to result from strong intermolecular interac- [44,46,71]. For example, poorly water-soluble curcumin has been co-
tions between them, it has been suggested that the dissolution rate of a amorphized with artemisin and administered to Sprague Dawley male
poorly soluble drug can be modified by the solubility/dissolution rate of rats for bioavailability analysis [44]. Because of the poor solubility of
the co-amorphous co-former [52,70]. When strong intermolecular in- pure crystalline coumarin, the drug was not detected in the plasma
teractions exist between the two components, the co-former might fa- when administered alone. On the contrary, a considerably high amount
cilitate the dissolution of a given poorly water-soluble drug. In other of coumarin was found when administering it for co-amorphous formu-
words, the dissolution rate of the poorly soluble component depends lation together with artemisin. However, no control groups, that is, pure
on the solubility of the co-former. In this regard, it was found that disso- amorphous coumarin or a physical mixture of coumarin and artemisin,
lution rates of several non-salt-forming co-amorphous indomethacin/ were analysed. Therefore, careful interpretation of these results must be
amino acid systems were dependant on the solubility of the co- exercised, as it is unclear whether the increase in bioavailability was
forming amino acid [51]. A similar approach was reported for ternary solely due to amorphization of coumarin or a result of its co-
co-amorphous naproxen/tryptophan/proline and naproxen/arginine/ amorphous formulation.

Please cite this article as: S.J. Dengale, et al., Recent advances in co-amorphous drug formulations, Adv. Drug Deliv. Rev. (2016), http://dx.doi.org/
10.1016/j.addr.2015.12.009
6 S.J. Dengale et al. / Advanced Drug Delivery Reviews xxx (2016) xxx–xxx

Table 3
Comparison of dissolution properties of co-amorphous systems.

Co-amorphous system Individual component Measured dissolution property Dissolution improvement Synchronized release Reference

Crystalline NAP 0.359 ± 0.016* NA


Crystalline CIM 0.669 ± 0.038* NA
Naproxen(NAP)/Cimitidine (CIM) Yes [57]
NAP co-milled 1.49 ± 0.108* 4-fold
CIM co-milled 1.32 ± 0.104* 2-fold
Crystalline IND 0.055 ± 0.002*
Crystalline NAP 0.300 ± 0.001*
Indomethacin (IND)/NAP Yes [36]
IND(NAP:IND 1:1) 0.491 ± 0.012* 9-fold
NAP (NAP:IND 1:1) 0.411 ± 0.003* 1.37-fold
Crystalline SVS 6.7 ± 1.4†
Crystalline GPZ 7.6 ± 2.9†
Simvastatin (SVS)/Glipizide (GPZ) NI [50]
GPZ (SVS:GPZ 1:1 BM) 28 ± 11† 3.7-fold
GPZ (SVS:GPZ 1:1 CM) 24 ± 15† 3.1-fold
Crystalline CBZ 0.034 ± 0.001* NA
CBZ (CBZ: Arginine:Tryptophan) 0.047 ± 0.005* Marginal (b0)
Carbamazepine (CBZ)/Amino acids NI [51]
CBZ (CBZ: Phenylalanine:Tryptophan) 0.041 ± 0.003* Marginal (b0)
CBZ (CBZ: Tryptophan) 0.037 ± 0.001* Marginal (b0)
IND (IND: Arginine: Phenylalanine) 11.19 * 200-fold
IND (IND: Arginine) 12.25 * 200-fold
IND/Amino acids NI [51]
IND (IND:Tryptophan) 0.096 * 1.5-fold
IND (IND:Phenylalanine) 0.175 * 3-fold
Crystalline RTV 0.13† NA
Crystalline IND NR NA
Ritonavir (RTV)/IND NI [45]
RTV (RTV:IND 1:1) 0.56† 4.3-fold
IND (RTV:IND 1:1) NR NR
Crystalline LH 0.0066* NA
Lurasidone HCl (LH)/Saccharine (SAC) NI [47]
LH (LH:SAC) 0.0371* 5.6-fold

NI Not investigated.
NR Not reported.
NA Not applicable.

Dissolution rates expressed in μg/ml.
⁎ Intrinsic dissolution rates expressed in mg cm−2 min−1.

In addition to poor solubility, poor intestinal permeability can add to a protection involves pumping the toxins back into the intestine. The
the low bioavailability of a given drug. Poorly soluble and poorly perme- process of protecting the body from toxins also extends to many
able drugs are classified as class 4 drugs in the biopharmaceutics classi- drugs, as they are also substrates of these efflux pumps. In order to over-
fication system (BCS) and are particularly problematic, because an come the challenges associated with poorly soluble drugs that are also
increase in dissolution and solubility only might not result in an in- efflux pump substrates, Teja et al. applied the co-amorphous formula-
creased bioavailability. In this context, poor permeability can be due to tion approach to address both problems simultaneously. In particular,
the physicochemical properties of the drug molecule, such as size and they co-amorphized the poorly soluble and permeable drug talinolol,
polarity, but often is due to being substrate of the so-called intestinal ef- an efflux pump substrate, with the efflux pump inhibitor naringin
flux pumps [77,78]. These efflux pumps are situated in the absorption [71]. Using this approach, they were also able to improve the dissolution
cell layer of the intestine and their main purpose is to prevent toxins rate of talinolol and simultaneously increase absorption of the drug
from entering the systemic blood circulation. The mechanism of such through efflux pump inhibition of the co-administered naringin. The
bioavailability of talinolol in wistar rats could thus be significantly im-
proved. The mean AUC(0 − t) from co-amorphous talinolol/naringin
was found to be 5.4-fold higher than the administration of pure crystal-
line talinolol (Fig. 3). In order to study the influence of naringin on the
permeability of talinolol, determination of permeability using the in
situ intestinal closed loop method is further studied. Permeability of
talinolol was found to slightly increase from 2.48 × 10− 5 (control
value without naringin) to 3.16 × 10−5 cm/s when formulated as co-
amorphous talinolol/naringin system. The increase in bioavailability of
talinolol was thus attributed to a combination of increased solubility
from the co-amorphous formulation and the inhibition of naringin by
efflux pump. Again, these findings have to be considered carefully as
no control group using pure amorphous talinolol or a co-
administration of physically mixed naringin was investigated. Thus, a
direct connection to the co-amorphous formulation approach cannot
be drawn.
The same group also investigated the bioavailability of co-
amorphous ritonavir/quercetin, where ritonavir is an efflux pump sub-
strate and quercetin is an efflux pump inhibitor [46]. However, unlike
talinolol/naringin, there was no statistically significant improvement
in the bioavailability (AUC) of ritonavir from its co-amorphous formula-
Fig. 1. Intrinsic dissolution profiles of co-amorphous lurasidone HCl–saccharin (LH–SAC,
tion compared with the crystalline control. Although a 5-fold improve-
triangles), pure amorphous LH (circles) and crystalline LH (squares) (adapted from ref. ment of the in vitro saturation solubility was achieved for ritonavir from
[47]). its co-amorphous system, the advantage could not be translated into a

Please cite this article as: S.J. Dengale, et al., Recent advances in co-amorphous drug formulations, Adv. Drug Deliv. Rev. (2016), http://dx.doi.org/
10.1016/j.addr.2015.12.009
S.J. Dengale et al. / Advanced Drug Delivery Reviews xxx (2016) xxx–xxx 7

Fig. 2. Powder dissolution profiles of crystalline, amorphous and co-amorphous (a) simvastatin (SVS) and (b) glibenclamide (GBC) formulations indicating long-lasting supersaturation
achieved using the co-amorphous drug–amino acid formulation approach. For clarifying the differences of the dissolution profiles of glibenclamide, a magnification of the first 90 min was
inserted. (Abbreviations: LYS = lysine, THR = threonine, SER = serine, CM = cryomilled, PM = physical mixture) (adapted from ref. [55]).

Please cite this article as: S.J. Dengale, et al., Recent advances in co-amorphous drug formulations, Adv. Drug Deliv. Rev. (2016), http://dx.doi.org/
10.1016/j.addr.2015.12.009
8 S.J. Dengale et al. / Advanced Drug Delivery Reviews xxx (2016) xxx–xxx

better conclusions on their behavior in vivo. For example, the co-


amorphous formulations need to be tested against the individual amor-
phous drugs as well as other competing technologies such as polymer-
and mesoporous silica-based glass solutions.
Down-streaming of the formulations into final dose forms, that is,
capsules or tablets, is another important factor that needs to be con-
sidered in future. For this purpose, more understanding on scalable
production techniques, such as spray-drying and melt-extrusion, is
necessary to make this approach feasible in an industrial set-up.
Furthermore, there is only little knowledge on the processability of
co-amorphous formulations. Amorphous materials have shown
problems with powder flow, pulverization and sticking to punches
in tablet presses [12]. Another issue could be to find a suitable gran-
ulation process. A wet granulation may not be feasible, as the intro-
duced water acts as a plasticizer and can induce crystallization. On
Fig. 3. Mean plasma concentration–time profile for pure talinolol (triangles) and co-
amorphous talinolol/naringin (squares) (adapted from ref. [71]).
the contrary, using melt granulation, the introduced heat could be
above the Tg of the co-amorphous formulation, which as well may
carry a risk for recrystallization. It has also been reported that amor-
significant in vivo outcome. On the contrary, the Tmax value was signifi- phous materials can have significant problems with disintegration
cantly reduced from 6 to 4 h for the co-amorphous formulation com- when formulated into tablets [79,80]. In this context, Lenz et al.
pared with pure crystalline ritonavir, indicating improved absorption. were the first to report a successful tablet formulation using the co-
This improvement was again attributed to enhanced solubility and per- amorphous approach [58].
meability due to co-amorphization and efflux pump inhibition. Howev- Despite the challenges, co-amorphous formulations have been
er, similar to the aforementioned studies, this study did not investigate shown to be a very promising and upcoming technique to stabilize the
any other controls. amorphous form of drugs, increase the dissolution and bioavailability
Overall, the few in vivo studies that have been performed using co- of poorly soluble drugs and overcome drawbacks related to other amor-
amorphous formulations indicated the potential of the co-amorphous phous formulation strategies.
approach. Nevertheless, there are still some unanswered questions
such as whether the improved bioavailability was due to pure
amorphization of the drug or a result of the co-amorphous formulation References
itself. Furthermore, a direct comparison with other amorphous stabili-
zation techniques such as polymer- or silica-based glass solutions [1] J. Aaltonen, T. Rades, Towards physico-relevant dissolution testing: the importance
of solid-state analysis in dissolution, Dissolut. Technol. 16 (2009) 47–54.
would be interesting. [2] B.C. Hancock, G. Zografi, Characteristics and significance of the amorphous state in
pharmaceutical systems, J. Pharm. Sci. 86 (1997) 1–12.
7. Concluding remarks and future outlook [3] W.L. Chiou, S. Riegelman, Pharmaceutical applications of solid dispersion systems, J.
Pharm. Sci. 60 (1971) 1281–1302.
[4] R. Laitinen, P.A. Priemel, S. Surwase, K. Graeser, C.J. Strachan, H. Grohganz, T. Rades,
The co-amorphous technology has established itself as a promising Theoretical considerations in developing amorphous solid dispersions, in: N. Shah,
approach to increase dissolution/solubility of poorly water-soluble H. Sandhu, D.S. Choi, H. Chokshi, A.W. Malick (Eds.), Amorphous Solid Dispersions,
Springer, New York 2014, pp. 35–90.
drugs and potentially to improve bioavailability. Although it has poten- [5] S.R.K. Vaka, M.M. Bommana, D. Desai, J. Djordjevic, W. Phuapradit, N. Shah, Excipi-
tial to become a platform technology to deal with these drugs, it is still a ents for amorphous solid dispersions, in: N. Shah, H. Sandhu, D.S. Choi, H. Chokshi,
new technology and the concept needs to be further established. A.W. Malick (Eds.), Amorphous Solid Dispersions, Springer, New York 2014,
pp. 123–161.
Most of the above-mentioned studies were conducted on the funda- [6] G. Van den Mooter, The use of amorphous solid dispersions: a formulation strategy
mental understanding of co-amorphous formulations, their stabiliza- to overcome poor solubility and dissolution rate, Drug Discovery Today: Technolo-
tion mechanism and their behavior during dissolution. However, there gies, 9 2012, pp. e79–e85.
[7] P.J. Marsac, S.L. Shamblin, L.S. Taylor, Theoretical and practical approaches for pre-
are still many unknowns and it is important to obtain a deeper under-
diction of drug–polymer miscibility and solubility, Pharm. Res. 23 (2006)
standing on the mechanisms of working of these systems. For example, 2417–2426.
similar to polymer-based glass solutions, the tendency of moisture up- [8] C. Leuner, J. Dressman, Improving drug solubility for oral delivery using solid disper-
take may also prove problematic for co-amorphous systems. Most of sions, Eur. J. Pharm. Biopharm. 50 (2000) 47–60.
[9] B.C. Hancock, S.L. Shamblin, G. Zografi, Molecular mobility of amorphous pharma-
the stability studies of co-amorphous mixtures have been conducted ceutical solids below their glass transition temperatures, Pharm. Res. 12 (1995)
in dry conditions. Thus, it would be interesting to study the effect of 799–806.
water on the performance of co-amorphous systems. [10] S. Janssens, G. Van den Mooter, Review: physical chemistry of solid dispersions, J.
Pharm. Pharmacol. 61 (2009) 1571–1586.
Particularly, in co-amorphous drug–excipient mixtures, there is no [11] Q. Lu, G. Zografi, Phase behavior of binary and ternary amorphous mixtures contain-
clear rationale on the suitability of an excipient as a co-former. In studies ing indomethacin, citric acid, and PVP, Pharm. Res. 15 (1998) 1202–1206.
using amino acids as excipients, it could be shown that certain amino [12] A.T.M. Serajuddin, Solid dispersion of poorly water-soluble drugs: early promises,
subsequent problems, and recent breakthroughs, J. Pharm. Sci. 88 (1999)
acids can be successfully co-amorphized with one drug but not neces- 1058–1066.
sarily with another. Therefore, finding a good co-former might be [13] T. Vasconcelos, B. Sarmento, P. Costa, Solid dispersions as strategy to improve oral
labor-intensive, given the plethora of different amino acids and other bioavailability of poor water soluble drugs, Drug Discov. Today 12 (2007)
1068–1075.
co-formers. [14] C. Brown, J. DiNunzio, M. Eglesia, S. Forster, M. Lamm, M. Lowinger, P. Marsac, C.
With respect to co-amorphous drug–drug formulations, it might be McKelvey, R. Meyer, L. Schenck, G. Terife, G. Troup, B. Smith-Goettler, C. Starbuck,
difficult to find a suitable partner molecule with a specific pharmacolog- Hot-melt extrusion for solid dispersions: composition and design considerations,
in: N. Shah, H. Sandhu, D.S. Choi, H. Chokshi, A.W. Malick (Eds.), Amorphous Solid
ical profile. The regulatory constraints and the compliance of adminis-
Dispersions, Springer, New York 2014, pp. 197–230.
tering more than one drug at the same time at a fixed ratio between [15] Y. Choudhari, H. Hoefer, C. Libanati, F. Monsuur, W. McCarthy, Mesoporous silica
the two drugs are other important issues that need to be addressed. drug delivery systems, in: N. Shah, H. Sandhu, D.S. Choi, H. Chokshi, A.W. Malick
The in vivo studies performed on co-amorphous drug formulations (Eds.), Amorphous Solid Dispersions, Springer, New York 2014, pp. 665–693.
[16] F. Wang, H. Hui, T.J. Barnes, C. Barnett, C.A. Prestidge, Oxidized mesoporous silicon
so far showed a potential in increasing the bioavailability of poorly sol- microparticles for improved oral delivery of poorly soluble drugs, Mol. Pharm. 7
uble drugs. However, further investigations are necessary to draw (2010) 227–236.

Please cite this article as: S.J. Dengale, et al., Recent advances in co-amorphous drug formulations, Adv. Drug Deliv. Rev. (2016), http://dx.doi.org/
10.1016/j.addr.2015.12.009
S.J. Dengale et al. / Advanced Drug Delivery Reviews xxx (2016) xxx–xxx 9

[17] S.W. Song, K. Hidajat, S. Kawi, Functionalized SBA-15 materials as carriers for con- [46] S.J. Dengale, S.S. Hussen, B.S.M. Krishna, P.B. Musmade, G. Gautham Shenoy, K. Bhat,
trolled drug delivery: influence of surface properties on matrix−drug interactions, Fabrication, solid state characterization and bioavailability assessment of stable bi-
Langmuir 21 (2005) 9568–9575. nary amorphous phases of ritonavir with quercetin, Eur. J. Pharm. Biopharm. 89
[18] K.K. Qian, R.H. Bogner, Application of mesoporous silicon dioxide and silicate in oral (2015) 329–338.
amorphous drug delivery system, J. Pharm. Sci. 101 (2011) 444–463. [47] S. Qian, W. Heng, Y. Wei, J. Zhang, Y. Gao, Coamorphous lurasidone hydrochloride–
[19] C.A. Prestidge, T.J. Barnes, C.-H. Lau, C. Barnett, A. Loni, L. Canham, Mesoporous silicon: saccharin with charge-assisted hydrogen bonding interaction shows improved
a platform for the delivery of therapeutics, Expert Opin. Drug Deliv. 4 (2007) 101–110. physical stability and enhanced dissolution with pH-independent solubility behav-
[20] R. Laitinen, K. Löbmann, C.J. Strachan, H. Grohganz, T. Rades, Emerging trends in the ior, Cryst. Growth Des. 15 (2015) 2920–2928.
stabilization of amorphous drugs, Int. J. Pharm. 453 (2013) 65–79. [48] M. Descamps, J.F. Willart, E. Dudognon, V. Caron, Transformation of pharmaceutical
[21] K. Löbmann, K.T. Jensen, R. Laitinen, T. Rades, C.J. Strachan, H. Grohganz, Stabilized compounds upon milling and comilling: the role of Tg, J. Pharm. Sci. 96 (2007)
amorphous solid dispersions with small molecule excipients, in: N. Shah, H. 1398–1407.
Sandhu, D.S. Choi, H. Chokshi, A.W. Malick (Eds.), Amorphous Solid Dispersions, [49] Y. Hu, K. Gniado, A. Erxleben, P. McArdle, Mechanochemical reaction of sulfathiazole
Springer, New York 2014, pp. 613–636. with carboxylic acids: formation of a cocrystal, a salt, and coamorphous solids, Cryst.
[22] N. Chieng, J. Aaltonen, D. Saville, T. Rades, Physical characterization and stability of Growth Des. 14 (2014) 803–813.
amorphous indomethacin and ranitidine hydrochloride binary systems prepared [50] K. Löbmann, C. Strachan, H. Grohganz, T. Rades, O. Korhonen, R. Laitinen, Co-
by mechanical activation, Eur. J. Pharm. Biopharm. 71 (2009) 47–54. amorphous simvastatin and glipizide combinations show improved physical stabil-
[23] S. Bates, G. Zografi, D. Engers, K. Morris, K. Crowley, A. Newman, Analysis of amor- ity without evidence of intermolecular interactions, Eur. J. Pharm. Biopharm. 81
phous and nanocrystalline solids from their X-ray diffraction patterns, Pharm. Res. (2012) 159–169.
23 (2006) 2333–2349. [51] K. Löbmann, H. Grohganz, R. Laitinen, C. Strachan, T. Rades, Amino acids as co-
[24] T. Rades, K.C. Gordon, K. Graeser, Molecular structure, properties, and states of mat- amorphous stabilizers for poorly water soluble drugs — part 1: preparation, stability
ter, in: L.V. Allen (Ed.), Remington — The Science and Practise of Pharmacy, Phar- and dissolution enhancement, Eur. J. Pharm. Biopharm. 85 (2013) 873–881.
maceutical Press, London, United Kingdom 2013, pp. 541–570. [52] K.T. Jensen, K. Löbmann, T. Rades, H. Grohganz, Improving co-amorphous drug for-
[25] D.Q.M. Craig, P.G. Royall, V.L. Kett, M.L. Hopton, The relevance of the amorphous mulations by the addition of the highly water soluble amino acid proline,
state to pharmaceutical dosage forms: glassy drugs and freeze dried systems, Int. Pharmaceutics 6 (2014) 416–435.
J. Pharm. 179 (1999) 179–207. [53] K. Löbmann, R. Laitinen, C. Strachan, T. Rades, H. Grohganz, Amino acids as co-
[26] H. Sandhu, N. Shah, H. Chokshi, A.W. Malick, Overview of amorphous solid disper- amorphous stabilizers for poorly water-soluble drugs — part 2: molecular interac-
sion technologies, in: N. Shah, H. Sandhu, D.S. Choi, H. Chokshi, A.W. Malick (Eds.), tions, Eur. J. Pharm. Biopharm. 85 (2013) 882–888.
Amorphous Solid Dispersions, Springer, New York 2014, pp. 91–122. [54] R. Laitinen, K. Löbmann, H. Grohganz, C. Strachan, T. Rades, Amino acids as co-amor-
[27] E. Fukuoka, M. Makita, S. Yamamura, Glassy state of pharmaceuticals. III.: thermal phous excipients for simvastatin and glibenclamide: physical properties and stabil-
properties and stability of glassy pharmaceuticals and their binary glass systems, ity, Mol. Pharm. 11 (2014) 2381–2389.
Chem. Pharm. Bull. 37 (1989) 1047–1050. [55] A.T. Heikkinen, L. DeClerck, K. Löbmann, H. Grohganz, T. Rades, R. Laitinen, Dissolu-
[28] L.M. Martínez, M. Videa, G.A. López-Silva, C.A. de los Reyes, J. Cruz-Angeles, N. tion properties of co-amorphous drug–amino acid formulations in buffer and
González, Stabilization of amorphous paracetamol based systems using traditional biorelevant media, Pharmazie 70 (2015) 452–457.
and novel strategies, Int. J. Pharm. 477 (2014) 294–305. [56] K.T. Jensen, F.H. Larsen, C. Cornett, K. Löbmann, H. Grohganz, T. Rades, Formation
[29] G.P. Johari, S. Kim, R.M. Shanker, Dielectric study of equimolar acetaminophen– mechanism of coamorphous drug–amino acid mixtures, Mol. Pharm. 12 (2015)
aspirin, acetaminophen–quinidine, and benzoic acid–progesterone molecular alloys 2484–2492.
in the glass and ultraviscous states and their relevance to solubility and stability, J. [57] M. Allesø, N. Chieng, S. Rehder, J. Rantanen, T. Rades, J. Aaltonen, Enhanced dissolu-
Pharm. Sci. 99 (2010) 1358–1374. tion rate and synchronized release of drugs in binary systems through formulation:
[30] P. Hoppu, K. Jouppila, J. Rantanen, S. Schantz, A.M. Juppo, Characterisation of blends amorphous naproxen–cimetidine mixtures prepared by mechanical activation, J.
of paracetamol and citric acid, J. Pharm. Pharmacol. 59 (2007) 373–381. Control. Release 136 (2009) 45–53.
[31] P. Hoppu, S. Hietala, S. Schantz, A.M. Juppo, Rheology and molecular mobility of [58] E. Lenz, K.T. Jensen, L.I. Blaabjerg, K. Knop, H. Grohganz, K. Löbmann, T. Rades, P.
amorphous blends of citric acid and paracetamol, Eur. J. Pharm. Biopharm. 71 Kleinebudde, Solid-state properties and dissolution behaviour of tablets containing
(2009) 55–63. co-amorphous indomethacin–arginine, Eur. J. Pharm. Biopharm. 96 (2015) 44–52.
[32] S. Schantz, P. Hoppu, A.M. Juppo, A solid-state NMR study of phase structure, molec- [59] S.L. Shamblin, L.S. Taylor, G. Zografi, Mixing behavior of colyophylized binary sys-
ular interactions, and mobility in blends of citric acid and paracetamol, J. Pharm. Sci. tems, J. Pharm. Sci. 87 (1998) 694–701.
98 (2009) 1862–1870. [60] P. Hoppu, A. Grönroos, S. Schantz, A.M. Juppo, New processing technique for viscous
[33] Y. Shimada, S. Goto, H. Uchiro, H. Hirabayashi, K. Yamaguchi, K. Hirota, H. Terada, amorphous materials and characterisation of their stickiness and deformability, Eur.
Features of heat-induced amorphous complex between indomethacin and lido- J. Pharm. Biopharm. 72 (2009) 183–188.
caine, Colloids Surf. B: Biointerfaces 102 (2013) 590–596. [61] P. Hoppu, J. Virpioja, S. Schantz, A.M. Juppo, Characterization of ultrasound
[34] Y. Shimada, S. Goto, H. Uchiro, K. Hirota, H. Terada, Characteristics of amorphous extrudated and cut citric acid/paracetamol blends, J. Pharm. Sci. 98 (2009)
complex formed between indomethacin and lidocaine hydrochloride, Colloids 2140–2148.
Surf. B: Biointerfaces 105 (2013) 98–105. [62] H. Wickström, M. Palo, K. Rijckaert, R. Kolakovic, J.O. Nyman, A. Määttänen, P.
[35] K. Pajula, L. Wittoek, V.-P. Lehto, J. Ketolainen, O. Korhonen, Phase separation in Ihalainen, J. Peltonen, N. Genina, T. de Beer, K. Löbmann, T. Rades, N. Sandler, Im-
coamorphous systems: in silico prediction and the experimental challenge of detec- provement of dissolution rate of indomethacin by inkjet printing, Eur. J. Pharm.
tion, Mol. Pharm. 11 (2014) 2271–2279. Sci. 75 (2015) 91–100.
[36] K. Löbmann, R. Laitinen, H. Grohganz, K.C. Gordon, C. Strachan, T. Rades, [63] P. Marsac, T. Li, L. Taylor, Estimation of drug–polymer miscibility and solubility in
Coamorphous drug systems: enhanced physical stability and dissolution rate of in- amorphous solid dispersions using experimentally determined interaction parame-
domethacin and naproxen, Mol. Pharm. 8 (2011) 1919–1928. ters, Pharm. Res. 26 (2009) 139–151.
[37] K. Löbmann, R. Laitinen, H. Grohganz, C. Strachan, T. Rades, K.C. Gordon, A theoret- [64] J.A. Baird, L.S. Taylor, Evaluation of amorphous solid dispersion properties using
ical and spectroscopic study of co-amorphous naproxen and indomethacin, Int. J. thermal analysis techniques, Adv. Drug Deliv. Rev. 64 (2012) 396–421.
Pharm. 453 (2013) 80–87. [65] K. Kawakami, Modification of physicochemical characteristics of active pharmaceu-
[38] S. Yamamura, H. Gotoh, Y. Sakamoto, Y. Momose, Physicochemical properties of tical ingredients and application of supersaturatable dosage forms for improving
amorphous salt of cimetidine and diflunisal system, Int. J. Pharm. 241 (2002) bioavailability of poorly absorbed drugs, Adv. Drug Deliv. Rev. 64 (2012) 480–495.
213–221. [66] L. Yu, Amorphous pharmaceutical solids: preparation, characterization and stabiliza-
[39] S. Yamamura, H. Gotoh, Y. Sakamoto, Y. Momose, Physicochemical properties of tion, Adv. Drug Deliv. Rev. 48 (2001) 27–42.
amorphous precipitates of cimetidine–indomethacin binary system, Eur. J. Pharm. [67] D.J. van Drooge, W.L.J. Hinrichs, M.R. Visser, H.W. Frijlink, Characterization of the
Biopharm. 49 (2000) 259–265. molecular distribution of drugs in glassy solid dispersions at the nano-meter scale,
[40] A. Shayanfar, A. Jouyban, Drug–drug coamorphous systems: characterization and using differential scanning calorimetry and gravimetric water vapour sorption tech-
physicochemical properties of Coamorphous atorvastatin with carvedilol and niques, Int. J. Pharm. 310 (2006) 220–229.
glibenclamide, J. Pharm. Innov. 8 (2013) 218–228. [68] M.M. Knopp, L. Tajber, Y. Tian, N.E. Olesen, D.S. Jones, A. Kozyra, K. Löbmann, K.
[41] T. Masuda, Y. Yoshihashi, E. Yonemochi, K. Fujii, H. Uekusa, K. Terada, Paluch, C.M. Brennan, R. Holm, A.M. Healy, G.P. Andrews, T. Rades, Comparative
Cocrystallization and amorphization induced by drug–excipient interaction im- study of different methods for the prediction of drug–polymer solubility, Mol.
proves the physical properties of acyclovir, Int. J. Pharm. 422 (2012) 160–169. Pharm. 12 (2015) 3408–3419.
[42] V. Tantishaiyakul, K. Suknuntha, V. Vao-Soongnern, Characterization of cimetidine– [69] S.L. Shamblin, X. Tang, L. Chang, B.C. Hancock, M.J. Pikal, Characterization of the time
piroxicam coprecipitate interaction using experimental studies and molecular dy- scales of molecular motion in pharmaceutically important glasses, J. Phys. Chem. B
namic simulations, AAPS PharmSciTech 11 (2010) 952–958. 103 (1999) 4113–4121.
[43] Y. Gao, J. Liao, X. Qi, J. Zhang, Coamorphous repaglinide–saccharin with enhanced [70] K.T. Jensen, L.I. Blaabjerg, E. Lenz, A. Bohr, H. Grohganz, P. Kleinebudde, T. Rades, K.
dissolution, Int. J. Pharm. 450 (2013) 290–295. Löbmann, Preparation and characterization of spray-dried co-amorphous drug–
[44] K. Suresh, M.K.C. Mannava, A. Nangia, A novel curcumin–artemisin coamorphous amino acid salts, J. Pharm. Pharmacol. (2015), http://dx.doi.org/10.1111/jphp.
solid: physical properties and pharmacokinetic profile, RCS Adv. 4 (2014) 12458.
58357–58361. [71] A. Teja, P.B. Musmade, A.B. Khade, S.J. Dengale, Simultaneous improvement of solu-
[45] S.J. Dengale, O.P. Ranjan, S.S. Hussen, B.S.M. Krishna, P.B. Musmade, G. Gautham bility and permeability by fabricating binary glassy materials of Talinolol with
Shenoy, K. Bhat, Preparation and characterization of co-amorphous ritonavir–indo- Naringin: Solid state characterization, in-vivo in-situ evaluation, Eur. J. Pharm. Sci.
methacin systems by solvent evaporation technique: improved dissolution behavior 78 (2015) 234–244.
and physical stability without evidence of intermolecular interactions, Eur. J. Pharm. [72] W. Brostow, R. Chiu, I.M. Kalogeras, A. Vassilikou-Dova, Prediction of glass transition
Sci. 62 (2014) 57–64. temperatures: binary blends and copolymers, Mater. Lett. 62 (2008) 3152–3155.

Please cite this article as: S.J. Dengale, et al., Recent advances in co-amorphous drug formulations, Adv. Drug Deliv. Rev. (2016), http://dx.doi.org/
10.1016/j.addr.2015.12.009
10 S.J. Dengale et al. / Advanced Drug Delivery Reviews xxx (2016) xxx–xxx

[73] C.T. Moynihan, Correlation between the width of the glass transition region and the [77] C.A. Lipinski, Drug-like properties and the causes of poor solubility and poor perme-
temperature dependence of the viscosity of high-Tg glasses, J. Am. Ceram. Soc. 76 ability, J. Pharmacol. Toxicol. Methods 44 (2000) 235–249.
(1993) 1081–1087. [78] D. Wagner, H. Spahn-Langguth, A. Hanafy, A. Koggel, P. Langguth, Intestinal drug ef-
[74] P.C. Painter, J.F. Graf, M.M. Coleman, Effect of hydrogen bonding on the enthalpy of flux: formulation and food effects, Adv. Drug Deliv. Rev. 50 (Suppl. 1) (2001)
mixing and the composition dependence of the glass transition temperature in S13–S31.
polymer blends, Macromolecules 24 (1991) 5630–5638. [79] J. Akbuga, A. Gürsoy, F. Yetimoĝlu, Preparation and properties of tablets prepared
[75] X. Lu, R.A. Weiss, Relationship between the glass transition temperature and the in- from furosemide-PVP solid dispersion systems, Drug Dev. Ind. Pharm. 14 (1988)
teraction parameter of miscible binary polymer blends, Macromolecules 25 (1992) 2091–2108.
3242–3246. [80] G. Owusu-Ababio, N.K. Ebube, R. Reams, M. Habib, Comparative dissolution studies
[76] D. Alonzo, G.Z. Zhang, D. Zhou, Y. Gao, L. Taylor, Understanding the behavior of for mefenamic acid-polyethylene glycol solid dispersion systems and tablets, Pharm.
amorphous pharmaceutical systems during dissolution, Pharm. Res. 27 (2010) Dev. Technol. 3 (1998) 405–412.
608–618.

Please cite this article as: S.J. Dengale, et al., Recent advances in co-amorphous drug formulations, Adv. Drug Deliv. Rev. (2016), http://dx.doi.org/
10.1016/j.addr.2015.12.009

Vous aimerez peut-être aussi