You are on page 1of 24

International Journal of Pharmaceutics 621 (2022) 121798

Contents lists available at ScienceDirect

International Journal of Pharmaceutics


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

Review

Solid-state and particle size control of pharmaceutical cocrystals using


atomization-based techniques
Aaron O’Sullivan , Barry Long , Vivek Verma , Kevin M. Ryan , Luis Padrela *
SSPC Research Centre, Department of Chemical Sciences, Bernal Institute, University of Limerick, Limerick, Ireland

A R T I C L E I N F O A B S T R A C T

Keywords: Poor bioavailability and aqueous solubility represent a major constraint during the development of new API
Supercritical fluids molecules and can influence the impact of new medicines or halt their approval to the market. Cocrystals offer a
Spray drying novel and competitive advantage over other conventional methods with respect towards the substantial
Nanoparticles
improvement in solubility profiles relative to the single-API crystals. Furthermore, the production of such coc­
Regulatory
rystals through atomization-based methods allow for greater control, with respect to particle size reduction, to
Antisolvent
Precipitation further increase the solubility of the API. Such atomization-based methods include supercritical fluid methods,
Multicomponent systems conventional spray drying and electrohydrodynamic atomization/electrospraying. The influence of process pa­
rameters such as solution flow rates, pressure and solution concentration, in controlling the solid-state and final
particle size are discussed in this review with respect to atomization-based methods. For the last decade, liter­
ature has been attempting to catch-up with new regulatory rulings regarding the classification of cocrystals, due
in part to data sparsity. In recent years, there has been an increase in cocrystal publications, specifically
employing atomization-based methods. This review considers the benefits to employing atomization-based
methods for the generation of pharmaceutical cocrystals, examines the most recent regulatory changes
regarding cocrystals and provides an outlook towards the future of this field.

1. Introduction utilized to improve API solubility (Desiraju, 2013). Crystal engineering


utilizes various predictive models to determine and predict crystal lat­
Some of the most significant obstacles facing the pharmaceutical tice structures. Cocrystal formation is one attractive crystal engineering
industry during the development of new APIs (Active Pharmaceutical approach for the improvement of API properties and inclusion of su­
Ingredients) include low aqueous solubility and thus, low oral pramolecular synthon prediction and cocrystallization theory are useful
bioavailability. It is reported that 40% of the top 200 marketed APIs crystal engineering tools (Desiraju, 2013). Cocrystals are an example of
have low solubility and up to 90% of drugs in the pipeline are estimated multicomponent crystalline materials which, unlike salts, do not require
to have the same issue, which is heavily influenced by the relatively ionizable groups, and unlike (co)amorphous solids, do not typically pose
recent increase in lipophilicity of new chemical entities (Sathisaran and issues with low thermodynamic instability (Bavishi and Borkhataria,
Dalvi, 2018; Rodriguez-Aller et al., 2015). The biopharmaceutical 2016). However, thermodynamically unstable cocrystals have been
classification system (BCS) divides APIs based on their solubility and identified, such as the Exemestane-maleic acid cocrystal (Jasani et al.,
permeability into four classes, with class II (low solubility, high 2019). Cocrystals are defined by the Food and Drug Administration
permeability) and class IV (low solubility, low permeability) drugs ex­ (FDA) as “solids that are crystalline materials composed of two or more
pected to account for a significant majority of new API molecules in the molecules in the same crystal lattice” (Aitipamula, 2012).
pipeline. Crystal engineering methods have been employed over the Conventional methods which have been employed for the production
years to overcome this issue. Crystal engineering utilizes a comprehen­ of pharmaceutical cocrystals include spontaneous contact formation,
sive knowledge of intermolecular interactions to predict and control liquid-assisted/neat grinding, extrusion-based methods, and solution
crystal packing and the resulting physical and chemical properties of the evaporation (Karimi-Jafari et al., 2018). Although these methods have
newly designed crystal structures, offering a set of methods which can be been successful in the formation of various cocrystals (Friscic and Jones,

* Corresponding author.
E-mail address: Luis.Padrela@ul.ie (L. Padrela).

https://doi.org/10.1016/j.ijpharm.2022.121798
Received 25 January 2022; Received in revised form 28 April 2022; Accepted 29 April 2022
Available online 4 May 2022
0378-5173/© 2022 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

2009; Weyna et al., 2009; Maheshwari et al., 2009), they provide little 2001; Narala, 2021). Single component crystals can be considered
control regarding the particle size of the resulting dried cocrystals. The anhydrous but may also exhibit polymorphism, where different molec­
inclusion of an atomization step within a cocrystallization procedure ular arrangements of the API can occur, which may display variations in
allows for more control on the resulting particle size, which has been stability, solubility and in other physicochemical properties (Higashi
shown to play a role in the solubility of drug molecules with a reduction et al., 2017). Multicomponent crystalline materials may be considered to
in particle size leading to an increase in solubility (Van Eerdenbrugh be in one of three major categories: hydrates/solvates, salts, and coc­
et al., 2010). Atomization involves the conversion of a bulk liquid into a rystals (Narala, 2021; Qiao et al., 2011). This review focuses on the
spray comprised of micro/nano-droplets containing the drug molecules. various atomization methods reported to date in the literature, which
This resulting spray can be dried far quicker than bulk solutions due to have demonstrated an ability to not only produce pharmaceutical coc­
the larger surface area. Particles also require a relatively short residence rystals, but control their resulting particle size and solid form, such as
time in the drying chamber after atomization due to this large surface hydrated and non-hydrated forms, as well as cocrystal polymorphs.
area, making this a suitable process for thermolabile compounds. A There has been debate as to the full description and definition of
significant risk in atomization-based methods relates to nozzle clogging, pharmaceutical cocrystals between researchers and regulatory bodies
which can occur with viscous liquids and when API molecules crystallize such as the US Food and Drug Administration (FDA). This debate will be
prior to atomization (Lefebvre and McDonell, 2017). Processes which covered in detail in section 4, but a recently adopted definition of coc­
utilize atomization during a cocrystallization process include spray rystals supplied by the FDA describes these as APIs themselves (rather
drying, electrospraying, spray congealing, and supercritical fluid (SCF) than intermediate products), allowing cocrystals to be considered as API
based methods such as supercritical enhanced atomization (SEA) and polymorphs or salts from a regulatory perspective. However, cocrystals
rapid expansion of supercritical solutions (RESS). themselves are still described in this guidance document as crystalline
Section 2 provides a comprehensive overview of pharmaceutical materials made up of two or more molecules within a single crystal
cocrystals including the mechanisms by which they are created, as well lattice. Unlike API polymorphs, cocrystals contain more than one type of
as their various advantages and disadvantages for pharmaceutical for­ molecule and, unlike salts, the cocrystal components interact non-
mulations. Following from this, atomization-based cocrystallization ionically (FDA, 2018). Cocrystals themselves can be further sub-
techniques are discussed in Section 3 based on their propensity to pro­ divided into categories based on the nature of their components. The
duce cocrystals, cocrystal solvates/hydrates and polymorphs. Section 4 most common type of cocrystals discussed in literature is molecular
discusses the regulatory aspects and outlooks for the commercial pro­ cocrystals, which are regarded as crystalline structures composed of at
duction and market accessibility of micro and nano-sized cocrystal least two neutral molecules (generally an API and coformer). Molecular
formulations. cocrystals are typically observed to maintain their cocrystalline struc­
ture by halogen or hydrogen bonds (Duggirala et al., 2016). Ionic coc­
2. Cocrystals rystals are another subtype of cocrystals which have been defined as
multicomponent species composed of a salt and an ionic or molecular
Within the pharmaceutical industry, APIs are known to exist in many compound. These cocrystals differs from salts as complete proton
different solid forms such as cocrystals, polymorphs, salts, and amor­ transfer is not observed in these intermolecular bonds (Smith et al.,
phous solid dispersions, with each possessing unique attributes to boost 2013). Although ionic cocrystals may occasionally be less thermody­
the physicochemical profile of the API. At a broad level, these APIs can namically stable than the corresponding single API, they generally
be classified as amorphous or crystalline, and an overview of these solid display stronger intermolecular bonds coupled with a greater degree of
forms is provided in Fig. 1. Amorphous solids may display a short-term change in their properties compared to the API (Duggirala et al., 2016;
molecular order but unlike crystalline solids, are incapable of devel­ Long, 2021; Braga et al., 2013).
oping any kind of long-term molecular order. Multicomponent amor­ Molecular and ionic cocrystals can be considered anhydrous coc­
phous materials include solvates/hydrates, salts and coamorphous rystals, cocrystal hydrates/solvates or cocrystals of salts (Qiao et al.,
materials (very weak to none intermolecular interactions) (Yu et al., 2011). Solvates are crystalline compounds, which include solvent

Fig. 1. Classification of the various solid state forms of APIs according to structure and composition of single/multicomponent forms (adapted from (Aitipa­
mula, 2012).

2
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

molecules in their crystal lattice in either a stoichiometric or non- at a macroscopic and microscopic view. Macroscopic level mechanisms
stoichiometric ratio (referred to as pseudopolymorphs by the FDA) tend to be determined based on the method of cocrystallization, and
(FDA, 2018). Solvates are referred to as hydrates when the solvent relate to bulk phase transformation events such as molecular diffusion,
involved is water (FDA, 2018). Aitipamula et al. stated that cocrystalline eutectic formation and amorphous phase-mediated cocrystallization
materials have a higher tendency to form hydrates/solvates than single (Friscic and Jones, 2009). Mechanisms of cocrystal formation at the
component crystals. This is based on their study highlighting a greater microscopic level are generally determined by the nature of the re­
percentage of cocrystal solvates/hydrates reported in Cambridge struc­ actants and the solvent. Such determinants include available functional
tural database (CSD) in comparison to single component crystalline groups, component solubilities and molecular ratio of API to the
hydrate/solvates (Aitipamula et al., 2010). However, Healy et al. stated coformer(s). It has been observed that certain cocrystals are only pro­
that cocrystallization can reduce the tendency of hydrate formation due duced in the presence of a specific solvent, for example the production of
to competitive hydrogen bonding of the coformer with the API, rather the 1:1 pimelic acid – 1,1-diferrocene cocrystal in the presence of
than with water molecules (Healy et al., 2017). methanol (Karimi-Jafari et al., 2018; Friscic and Jones, 2009). Micro­
Similarly to API crystals, cocrystals can also exhibit polymorphism scopic cocrystal formation involve molecular recognition events such as
which may result in cocrystal polymorphs with altered physiochemical the formation of supramolecular synthons and hydrogen and /or
properties (Qiao et al., 2011). It is estimated that up to 50% of APIs halogen bonds (Friscic and Jones, 2009).
exhibit at least one polymorphic form and cocrystals can also fall under The various design-based approaches to determine the possibility of
this category (Aitipamula et al., 2010). Polymorphs form due to struc­ producing a particular cocrystal or for the screening of cocrystals
tural changes of cocrystal components/reactants, but changes leading to include the use of: CSD, structural analysis for hydrogen bond pro­
polymorphs are also seen within the intermolecular interactions be­ pensity, supramolecular synthon prediction, Hansen solubility parame­
tween these components (Qiao et al., 2011). As such, there are several ters, and pKa values (Buddhadev and Garala, 2021; Salem et al., 2019).
major distinctions of cocrystal polymorphs such as conformational Bonding between API molecules and coformers (or other APIs) is pri­
polymorphs, synthon polymorphs and tautomeric polymorphs. Synthon marily by non-covalent interactions such as hydrogen and halogen
polymorphs, generally refer to alterations within the hydrogen bonding bonding as mentioned earlier, and Van Der Waals attractions. Although
motif, or synthons. These supramolecular synthons are structural crystal cocrystal prediction is still in its early years, methods such as empirical
units formed by intermolecular interaction and can be considered homo- observation and an analysis of the available hydrogen bond donor and
or hetero-synthons (Reddy et al., 1996). Homosynthons are formed as acceptors provide an alternative route for cocrystal prediction and for­
dimers by two similar reacting functional groups, whereas hetero­ mation, limiting the amount of ‘trial and error’ experiments required.
synthons are formed by different reacting functional groups. In many However, the most successful method for simple design based cocrystal
cases, this form of polymorphism causes a change from a cocrystal prediction is the consideration of supramolecular synthons (Buddhadev
containing both a homosynthon and a heterosynthon, to a polymorph and Garala, 2021). Intermolecular interactions allow for the formation
containing two heterosynthons (Mukherjee and Desiraju, 2011; Aitipa­ of structurally significant crystal units which are known as supramo­
mula et al., 2014). One example of this are the polymorphs observed lecular synthons (Reddy et al., 1996). In recent years, this synthon-based
within the 4-hydroxybenzoic acid and 4,4-bipyridine cocrystal. Form I design coupled with the propensity for both hydrogen and halogen
contains a hydroxyl-pyridine synthon and an acid dimer, whereas the bonding has allowed for a much more reliable medium to predict coc­
form II polymorph contains two heterosynthons, an acid-pyridine syn­ rystal formation. Common reacting functional groups involved in su­
thon and a carbonyl-hydroxyl synthon. It must also be noted that this pramolecular synthons during cocrystal formation include carboxylic
form of polymorphism is more prevalent in cocrystal species with a high acids, alcohols, and amides (Buddhadev and Garala, 2021). It has also
number of hydrogen bonds or the capacity for a high number of been observed that there is a far higher propensity for the formation of
hydrogen bonds (Mukherjee and Desiraju, 2011; Aitipamula et al., heterosynthons in comparison to homosynthons. A CSD analysis illus­
2014). Another classification of cocrystal polymorphs are conforma­ trated that 34% of molecular carboxylic acid entries form homosynthons
tional polymorphs. This refers to the ability of ‘flexible’ molecules to whereas the remaining 66% from heterosynthons. Similarly, only 26%
exhibit different molecular conformations within the molecules itself (or of the molecular alcohol entries have been observed forming homo­
within the coformer), but such changes may also be seen within inter­ synthons whereas the remaining 74% form heterosynthons (Shattock,
molecular bonds, as seen in cocrystals. It is this flexible nature seen in 2008). An analysis using the CSD specifically highlighted that alcohol-
many API molecules which allow for the high occurrence of API poly­ alcohol and COOH-COOH supramolecular homosynthons are less fav­
morphism. This polymorphism is primarily seen in cocrystal structures oured than supramolecular heterosynthons alcohol-aromatic nitrogen
which contain a molecule with degrees of torsional freedom allowing for and acid-aromatic nitrogen. However, when both supramolecular het­
molecular structure alterations (Aitipamula et al., 2014). Packing erosynthons may be formed, the acid-aromatic nitrogen is more strongly
polymorphs are another form of cocrystal polymorphism and refers to favoured (Shattock, 2008).
changes in the overall three-dimensional packing of cocrystals. There is The propensity of whether two reactants form a cocrystal or a salt
no clear indication as to when this form of polymorphism will occur, but may be determined based on pKa values. If the difference of pKa values
unlike conformational polymorphs, packing polymorphs generally occur (base – acid) (ΔpKa) is greater than three, it is likely that a salt will form
in cocrystal systems with limited conformational freedom. The final type under appropriate conditions. Contrary to this, a ΔpKa value less than
of cocrystal polymorphism is tautomeric polymorphs (Aitipamula et al., zero is a good indication that a cocrystal will form rather than a salt.
2014). These polymorphic forms arise due to changes in the position of Some issues may reside in ΔpKa values between zero and three, which
protons and/or electrons and occurs due to a proton moving from one has been described as a cocrystal to salt continuum, where the pro­
polar atom to another allowing the transformation of hydrogen donor pensity for either a cocrystal or salt to form is unclear and difficult to
atom to a hydrogen acceptor, and the inverse happens to another atom determine (Buddhadev and Garala, 2021). As mentioned earlier, sol­
(Martin, 2009). The formation of these various multicomponent crys­ vates and hydrates are another form of multicomponent crystalline
talline materials happens by means of several mechanisms, giving rise to materials that can be composed of a single molecule with incorporated
a defined and repeating crystal lattice structure. solvent molecules, or a cocrystal system may incorporate a solvent
molecule within its crystal lattice also. This may be formed using
2.1. Mechanisms of cocrystallization solvent-assisted cocrystallization methods such as wet grinding, solvent
evaporation, and spray drying, where the solvent utilized for cocrys­
When discussing the mechanism behind the formation of multi­ tallization is entrapped within the cocrystal structure, or by dry/neat
component solids such as cocrystals, it is necessary to look at the solids methods such as neat grinding in which one or more of the reactants is a

3
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

solvate/hydrate. This allows the solvent molecules within the solvated ionizable groups/sites. The generation of amorphous compounds rather
reactant(s) to be included within the resulting cocrystal solvate structure than crystalline has also shown a positive influence on compound sol­
(Friscic and Jones, 2009). ubility, however thermodynamic instability issues may arise which can
cause conversion to a more stable crystalline form during manufacture,
storage or transportation. Consequently, significant testing by regula­
2.2. Impact of cocrystallization on API physicochemical properties
tory bodies is needed to ensure that no conversion occurs before being
administered to patients (Bavishi and Borkhataria, 2016). Amorphous
Multicomponent crystalline materials as discussed above, provide an
solid dispersions (ASD) have proven quite effective in enhancing the
avenue by which pharmaceutical companies can improve API properties
solubility and bioavailability of pharmaceutical formulations, however,
such as solubility, bioavailability, and permeability. However, these
there is a lack of data explaining the underlying mechanisms behind
improved properties, relative to the single component API cannot be
drug liberation and permeation from ASDs (Schittny et al., 2020).
estimated or predicted easily and must be experimentally characterized
Cocrystals however can be produced regardless of whether the reactants
during the generation of a new multicomponent system. Prediction of
contain ionizable groups, and cocrystals have exhibited higher ther­
cocrystal properties has only been successful by a small number of
modynamic stability. Cocrystals utilize supersaturation to improve their
methods, such as the construction of artificial neural networks (Gamidi
aqueous solubility through the creation of metastable and supersatu­
and Rasmuson, 2020). These altered properties can have significantly
rated states. This overall mechanism has been described as the spring
beneficial effects such as improved solubility, dissolution rate and
and parachute phenomenon (Bavishi and Borkhataria, 2016). As coc­
bioavailability but can also have adverse effects such as increased
rystals are usually held together by relatively weak hydrogen bond(s),
toxicity which may require significant testing to be performed prior to
dissociation within an aqueous biological system is relatively rapid. The
filing to patent a new cocrystal system.
more soluble coformer solubilizes within the aqueous medium leaving
the hydrophobic API molecules behind which are supersaturated. This
2.2.1. Solubility, dissolution rate and bioavailability
high energy drug form immediately precipitates and is known as the
Perhaps the most valuable property of a cocrystal is an increase in its
spring (within the spring and parachute phenomenon). This creates
aqueous solubility. A major method for solubility enhancement in the
mildly aggregated API clusters. But to maintain this newly supersatu­
past and at present in certain cases, is salt formation (Bavishi and Bor­
rated solution to allow for adsorption, precipitation inhibitors
khataria, 2016). However, salt formation is limited to APIs with suitable

Table 1
Summary of reported/produced pharmaceutical cocrystals during the year 2021, and the associated increase in apparent solubility compared to the raw API.
API Coformer Ratio Method Increase in Solubility Relative to Reference
Raw API

Artemisinin Orcinol 1:1 Solvent evaporation 26-fold (Kaur et al., 2021)


Resorcinol 2:1 Seeded solvent evaporation 21-fold
Carvedilol Tartaric acid – Antisolvent precipitation 2000-fold (Mohammady et al., 2021)
ultrasonication
Rosuvastatin L-Asparagine 1:1 Solvent evaporation 2.17-fold (Vemuri and Lankalapalli,
L-Glutamine 1:1 1.60-fold 2021)
Metaxalone Nicotinamide 1:1 Solvent evaporation 8.6-fold (Gohel et al., 2021)
4-hydroxybenzoic acid 1:1 3.4-fold
Diltiazem Fumaric acid 1:1 Solvent Evaporation − 16.5-fold (Diniz, 2021)
Aceclofenac Salicylic acid 1:1 Neat grinding 7-fold (Pekamwar and Kulkarni,
2021)
Azithromycin Neotame 1:1 Solvent Evaporation < 2-fold (Upadhye, xxxx)
loratadine Succinic acid 1:1 Solution method 2-fold (Setyawan et al., 20212021)
5-Flurouracil phenolic acid nutraceutical ferulic acid 1:1:1 Liquid assisted grinding 1.64-fold (Yu et al., 2021)
+ water
Ceritinib Nicotinamide 1:1 Neat grinding 119-fold (Awasthi et al., 2021)
1:2 155.3-fold
2:1 83.1-fold
Diacerein β-Resorcylic Acid 1:3 Antisolvent crystallization 2.3 – 2.8-fold (Patel and Raval, 2020)
Apixaban Caffeine 1:1 – 3-fold* (Madan et al., 2021)
Paliperidone P-hydroxybenzoic acid 1:1 Solvent evaporation 141.6-fold (Thimmasetty et al., 2021)
Aceclofenac Dimethyl urea 1:2 Liquid/neat grinding 3.43-fold* (Afzal et al., 2021)
Curcumin Ascorbic acid – Solvent evaporation 576-fold (Pantwalawalkar et al.,
2021)
Naproxen L-alanine 1:1 Liquid assisted grinding 1.4-fold* (Latif, 2021)
Rebamipide Oxalic acid 1:1 Liquid assisted grinding 7.29-fold (Jindal et al., 2021)
Citric acid 12.58-fold
3,4-dihydroxybenzoic acid 11.36-fold
Nevirapine Salicylamide – Solvent drop grinding 4.5-fold (Panzade et al., 2021)
3-hydroxybenzoic acid
Ethionamide Baicalein 1:1 Seeded solvent evaporation 8.71-fold (Wang, 2021)
Axitinib Fumaric acid 1:1.5 Liquid assisted grinding + slurrying 11.7-fold (Ren et al., 2021)
Suberic acid 1:1 5.5-fold
Trans-cinnamic acid 1:1 1.9-fold
Bergenin Isonicotinamide 1:1 Liquid assisted grinding + rapid 1.57 – 2.87-fold (Liu, 2021)
solvent removal
Telmisartan 3,5-dihydroxy benzoic acid 1:1 Liquid assisted grinding + slurrying 3-fold (Yu, 2021)
Nicorandil Suberic acid 1:0.5 Liquid assisted grinding 1.9-fold (Mannava et al., 2021)
Lornoxicam 1,3-dimethyl urea 1:3 Liquid assisted grinding, solvent 2.52-fold (Fatima et al., 2021)
evaporation
Telmisartan Oxalic acid – – 7-fold (Dhibar et al., 2021)

* fold increase related to the area under the curve (AUC) rather than solubility.

4
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

(parachute) are incorporated. Over time, this saturated amorphous between the types of supramolecular synthons and the resulting coc­
phase solution (which has been given sufficient time to reach a highly rystals melting point, it has been observed that packing efficiency plays
soluble dose) will convert to a metastable form with a superior solubility a larger role in determining the melting point and thus, the thermal
before converting to the final, stable polymorph (Bavishi and Borkha­ stability of the resulting cocrysnotal. A study by Wang et al. looked at the
taria, 2016). There have been several reported cases in which solubility melting points of three cocrystals of the same API, apremilast, a BCS
has been improved for poorly soluble drugs cocrystallization such as class IV drug used to treat psoriasis, with three different coformers:
Ritonavir cocrystallized with adipic acid (6-fold increase), and dip­ nicotinamide, caffeine, and acetylsalicylic acid. It was determined that
fluzine cocrystallized with benzoic acid (500-fold increase) (Bavishi and the packing efficiency for the nicotinamide and caffeine cocrystals were
Borkhataria, 2016; Lin et al., 2014). Table 1 summarizes several recently 71.3 % and 75.6 % which is believed to be the reason behind the higher
reported pharmaceutical cocrystals and the increase in API solubility melting point of the apremilast-caffeine cocrystal in comparison to the
seen in comparison to the pure API. Kuminek et al. cocrystallized the apremilast-nicotinamide cocrystal. The higher melting point of caffeine
BCS class II drug Posaconazole with p-aminobenzoic acid with a stoi­ over nicotinamide is also believed to influence this. An increase in
chiometric ratio of 2:3. The solubilities and dissolution rates for this packing efficiency is indicative of strong structural interactions. These
cocrystal far exceeded those of the pure Posaconazole drug substance cocrystals exhibited lower melting temperatures than the API but higher
during studies carried out in a medium which mimicked physiological than each of the coformers. The apremilast-acetylsalicylic acid cocrystal
conditions within intestinal fluids (Kuminek et al., 2019). displayed a greater packing efficiency and in turn, also displayed greater
Conversely, solubility has also been observed to decrease upon coc­ thermal stability seen by the cocrystals melting point of 183.5 ◦ C and
rystallization. For many pharmaceutical compounds (BCS Class II and decomposition temperature of 180.4 ◦ C (Wang et al., 2018). This is an
IV), this may be considered to be a major disadvantage to the majority of excellent example of the relationship between packing efficiency and
poorly soluble compounds but can serve to benefit when a prolonged resulting thermal stability.
residence time within the body is required. One such example of this was Chemical stability may also become altered due to cocrystallization
achieved with a broad-spectrum antibiotic used to treat tuberculosis, of an API with an appropriate coformer. Chemical stability is regarded as
moxifloxacin. Eedara et al. produced a new unreported cocrystal of any structural alterations of the API cocrystal by chemical reaction. Such
moxifloxacin with cinnamic acid and demonstrated that an API to reactions include oxidation and reduction reactions, as well as hydro­
coformer ratio of 1:2 produced a cocrystal with a decreased solubility lysis and dehydration reactions, but few studies exist in which chemical
and slower dissolution rate, which is hypothesised to increase the resi­ stability is directly accessed upon cocrystallization. Guo et al. investi­
dence time within the lung of tuberculosis patients (Eedara et al., 2018). gated improving the stability of the drug nicorandil, which is generally
This is one of the few studies which reported a decrease in solubility of a regarded as chemically unstable, as well as being unstable at higher
cocrystal in comparison to the stable API molecule. However, despite the temperatures and higher humidity. Cocrystallization of this drug by
limited reports of reduced solubility, this system highlights that coc­ reaction cocrystallization was carried out with various coformers,
rystals may also be used when time-delayed release of an API is neces­ namely salicylic acid, 2,5-dihydroxybenzoic acid, 3-hydroxybenzoic
sary, although increases in solubility are usually the aim with cocrystals. acid, and 1-hydroxy-2-naphthoic acid. The lack of stability of the pure
Oral bioavailability of a drug refers to the amount of the API that is drug nicorandil can be attributed to self-catalysed decomposition due to
available to enter the systemic circulatory system and low oral crystal packing. However, the addition of such organic acid coformers
bioavailability is a major constraint within the pharmaceutical industry within the crystal lattice allowed for the realignment of the nicorandil
(Kumar et al., 2018). The use of polymers and excipients to maintain crystals preventing this form of decomposition. Hydrothermal stability
supersaturation and delay precipitation (parachutes within the spring was also increased due to cocrystallization as long-term storage at 40 ◦ C
and parachute phenomenon], have been shown to positively influence and a relative humidity of 75 % caused a large reduction of stable pure
the bioavailability of cocrystalline APIs (Karimi-Jafari et al., 2018). nicorandil with a final content of only 2 % after 20 days of storage. The
Dipfluzine-benzoic acid cocrystals showed a significant improvement cocrystals, however, retained a nicorandil content greater than 90 %
over the pure drug substance in all previously mentioned areas of sol­ under the same storage conditions demonstrating their superior stability
ubility, dissolution rate and bioavailability and is an excellent example over the pure drug. Among these improvements, other improvements
of the extent of which cocrystallization with an appropriate coformer seen within the cocrystal include an increase in thermal stability and
can have on resulting drug properties (Lin et al., 2014). Martin et al. dissolution rates in comparison to the pure drug substance (Guo et al.,
produced a ketoconazole – p-aminobenzoic acid cocrystal for the 2020). This is an excellent example as to the many benefits a cocrystal
enhancement of drug properties, of this BCS class II drug. The resulting can induce on a drug system when appropriate coformers and produc­
cocrystal demonstrated an aqueous solubility ten times greater than the tion methods are utilized.
pure drug substance along with almost a 7-fold increase in bioavail­ Other forms of stability such as photostability have been seen to vary
ability. Additionally, the antifungal effect of this drug was significantly after cocrystallization, although it has not been reported extensively in
increased after cocrystallization with p-aminobenzoic acid (Martin the literature. Photostability refers to a substance propensity to react
et al., 2020). during prolonged exposure to light. Cocrystallization has been reported
to influence the photostability of APIs. Shinozaki et al. produced a 1:1
2.2.2. Stability cocrystal from the API Levofloxacin with another API, Metacetamol. The
Determining the stability of pharmaceutical cocrystals relative to resulting cocrystal proved to have a reduced photodegradation than the
other crystalline components such as solvates, and polymorphic forms of corresponding Levofloxacin hydrate indicating the ability of cocrystal­
the APIs is a critical requirement during the filing of a new chemical lization to alter the photostability of an API molecule (Shinozaki et al.,
entity. Such studies include thermal stability, chemical stability, solu­ 2019). This may lengthen the drugs storage light or allow it to be stored
tion stability and relative humidity stability during the manufacture, in an area with light exposure, without concerns regarding losses in drug
transportation and storage stages (Kumar et al., 2018). Thermal stability stability. A similar instance occurred during the cocrystallization of
analysis can be carried out by hot stage microscopy to determine the furosemide with various coformers. Upon cocrystallization, it was
phase transition conditions, as well as differential scanning calorimetry observed that the absorption range of the furosemide-nicotinamide
(DSC) which is primarily utilized to determine a substances melting cocrystal shifted to a shorter wavelength in comparison to pure furose­
point. The melting point of a substance provides a strong indication of mide and the other cocrystals. This allowed for an increase in photo­
thermal stability, with an increase in melting point of the cocrystal stability for the resulting furosemide cocrystal.
relative to the API suggesting an increase in thermal stability (Kumar
et al., 2018; Shaikh et al., 2018). Although there may be a minor link

5
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

2.2.3. Permeability tion rate (Sun et al., 2012; Hanada et al., 2014). The Noyes Whitney (1)
Poor permeability is observed in BSC class III and IV drugs. Perme­ equation states that the dissolution rate increases with a reduction in
ability refers to a drugs ability to cross biological membranes to enter particle size. In this equation, dC/dT refers to the dissolution rate, D is
areas of the body, most notably the circulatory system through the in­ the drugs diffusion coefficient, A is the drug particles effective surface
testinal wall (Lennernaäs, 1998). A link between solubility and perme­ area, h is the thickness of the diffusion membrane, Cs refers to the drugs
ability has been observed when permeability is driven by passive saturation solubility at the diffusion layer and Cb refers to the drug
diffusion, which is directly related to the concentration gradient (Rao concentration (Mathur et al., 2011; Dizaj et al., 2015). Wang et al.
et al., 2011; Zhang et al., 2019). An increase in solubility leads to a investigated the particle size reduction of coenzyme Q10 in an attempt
greater concentration gradient, leading to increased permeability. Other to improve dissolution rate, without the use of stabilizers, additives and/
factors play a role in this also such as particle size and polarity (Rao or surfactants and showed an increase in apparent solubility, dissolution
et al., 2011; Zhang et al., 2019). Hydrochlorothiazide (HCZ) has been rate and bioavailability as particle size decreased (Sun et al., 2012).
extensively studied as a cocrystallization candidate (Sanphui et al., There has been an increase in recent years in the number of papers
2015; Ranjan et al., 2017; El-Gizawy et al., 2015; Gopi et al., 2017; Al- published in the literature on nano-sized cocrystals, and a variety of
Otaibi et al., 2019; Al-Otaibi et al., 2020), as it is a BCS class IV drug methods have been utilized for this (Mohammady et al., 2021; Tan et al.,
exhibiting low solubilities and permeabilities. Sanphui et al. cocrystal­ 2021; Karashima et al., 2016; Pi et al., 2019; Bhandari et al., 2020;
lized HCZ with the coformers nicotinic acid, nicotinamide, p-amino­ Salem et al., 2021; Witika et al., 2021; Cocrystals, 2021).
benzoic acid, succinimide and resorcinol via the wet grinding method
dC DA(CS − Cb )
(Sanphui et al., 2015). Solubilities were increased for all cocrystals in = (1)
dt h
comparison to the pure drug substance except for cocrystals incorpo­
rating nicotinic acid and succinimide, which displayed reduced aqueous Sander et al. produced a caffeine and dihydroxybenzoic acid nano-
solubilities. Permeability of the API and the cocrystals were investigated cocrystal using the sonochemistry method (Sander et al., 2010).
using Franz diffusion cells with the help of a MW 14000 Da dialysis Initially, the API and coformer were separately dissolved in acetone and
membrane. The cocrystal which utilized succinimide as coformer had a injected into an antisolvent (hexane) while simultaneously being soni­
reduced permeability which may be attributed to its reduced solubility. cated to allow for nanocrystallization, with cocrystal particles collected
Cocrystals of nicotinamide and nicotinic acid displayed the highest by filtration. Scanning Electron Microscopy (SEM) confirmed the pres­
solubilities. These cocrystals not only displayed a higher initial perme­ ence of plate-shaped particles of the cocrystal with a large particle size
ability, but also displayed a superior permeability once a steady state distribution, ranging from 200 nm to 5 μm. To produce cocrystal par­
was reached, as determined by the concentration gradient (Sanphui ticles with a reduced particle size, a two solvent system was then carried
et al., 2015). Another BSC class IV drug, acetazolamide, was investigated out by choosing two solvents (chloroform and acetone) with which the
to determine if its low solubility and permeability could be improved by respective solutes have a higher solubility, promoting rapid nucleation
forming multicomponent crystalline materials. A cocrystal of acetazol­ (Sander et al., 2010). When a surfactant was included, the resulting
amide with theophylline was formed by liquid assisted grinding fol­ cocrystals displayed a particle size 136.4 nm with reduced agglomera­
lowed by solvent evaporation. In addition, a salt of acetazolamide with tion (Sander et al., 2010). Hickey et. al. (Hickey et al., 2007) investi­
piperazine was formed by reaction crystallization. Salt formation was gated the control of particle size of a carbamazepine cocrystal, with
predicted as ΔpKa was found to be 2.35. Although this lies within the saccharin as conformer, and compared the resulting cocrystal with the
cocrystal to salt continuum, the authors understood that this was a good marketed drug substance. The cocrystal was prepared by conventional
indication of salt formation as ΔpKa was close to 3.0 (Buddhadev and cooling precipitation from an alcohol. Metal mesh sieves were used for
Garala, 2021; Zhang et al., 2019). The generation of infrared spectra the separation of cocrystal particles based on crystal size. As expected, a
confirmed the presence of the cocrystal and salt. Solubility studies reduction in particle size generally led to an increase in solubility, with
showed an increase in solubility at higher pH values for the cocrystal, the largest mean dissolution and largest total percentage dissolved after
due to its slight acidity. The opposite was seen for the salt as it is slightly 60 min seen in the sample with the smallest size particles (<53 μm)
alkaline, and its solubility decreased with increasing pH solutions. At pH (Hickey et al., 2007). These studies have demonstrated an increase in
6.8, the concentrations of the API acetazolamide solution, the cocrystal cocrystal solubility through particle size reduction. Mechanochemistry
solution and the salt solution were 0.72, 2.18 and 10.59 mg/ml, methods such as neat and liquid assisted grinding (LAG) have been used
respectively. The difference between the solubilities of the salt and for the production of various types of cocrystals (also in the nano-size
cocrystal can be attributed to the higher solubility of the coformer range) but are incapable of direct control over resulting particle size,
piperazine (for the salt) compared to theophylline (for the cocrystal), and undesirable solvates may also occur. Polymer assisted grinding
and to the lower melting point of the salt relatively to the cocrystal (POLAG) overcomes the solvate issues by employing a solid or liquid
(confirmed by TG-DSC). Permeation diffusion of the salt and cocrystal polymer which is incapable of forming a solvate (Hasa et al., 2015).
were also investigated, using a Franz diffusion cell at pH 6.8. Both the There are a substantial number of top-down and bottom-up methods
salt and the cocrystal showed a larger degree of permeability during the which appear to provide varying levels of control over the final particle
initial lag phase and once a steady state was reached in comparison with size of pharmaceutical materials. One of the most popular top-down
the pure drug substance. The cocrystal performed better than the salt methods is milling which allows for the reduction in size of previously
despite the higher solubility of the salt. This is believed to be due to the produced particles through particle attrition. Fully stable particles are
increased polarity of the salt negatively effecting its diffusion across the usually dry milled while particles with a tendency to agglomerate can be
membrane (Zhang et al., 2019). Production of cocrystals and even salts milled in liquid suspension with stabilizing surfactants. Wet milling
of poorly soluble and/or poorly permeable API molecules has been usually produces smaller particles with a narrower size distribution.
shown to be an effective and flexible method to improve various prop­ Unfortunately, sieving is required to achieve a desired particle size
erties of pharmaceutical drugs. distribution, and due to the low flowability and cohesive nature of the
resulting particles, the addition of excipients is generally required. Other
2.3. Particle size control issues with this process include a high energy consumption and the in­
clusion of undesirable mill surface materials (Kumar et al., 2021). High
The topic of particle size control within the pharmaceutical industry, pressure homogenisation (HPH) is another method for particle size
in particular small molecule APIs, generally aims at reducing particle reduction which involves the passing of a drug suspension through a
size and reducing particle size distribution in an effort to improve and small aperture<20 µm. This high shear method causes the disruption of
control drug physicochemical properties such as solubility and dissolu­ particles suspended in a non-solvent leading to the production of smaller

6
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

particles. The movement of particles from a high pressure zone to a low 2015, is the first and only commercially available multidrug cocrystal in
pressure zone also leads to particle size reduction through cavitation that both reactants, Sacubitril and Valsartan, are API molecules. Sacu­
(Kumar et al., 2021; Preiss et al., 2022). Although this method prevents bitril improves the neurohormonal system within the heart whilst val­
drug amorphization and polymorphic conversion and has a lower me­ sartan reduces the damage caused by angiotensin II to allow for effective
chanical energy requirement than milling. However, clogging can occur treatment of heart failure (Feng et al., 2012; Fala, 2015). Odomzo is a
without a preliminary particle size reduction step prior to this method diphosphate molecule formed through the bonding of a monophosphate
(Kumar et al., 2021). Atomization based processes are attractive method salt (sonidegib monophosphate) with phosphoric acid and acts as a cell
to control and reduce particle sizes. These methods offer significant surface kinase inhibitor. This cocrystal was approved in 2015 for the
flexibility since the resulting particle size is influenced by the chosen treatment of basal cell carcinomas (Roy and Ghosh, 2020). Steglato is a
atomization/drying method and through altering process parameters cocrystal of etugliflozin and L-pyroglutamic acid and was produced
and set-up configurations (Kumar et al., 2021). through a collaborative effort by Pfizer and MSD. Similarly to Suglat, it
is a sodium glucose transport 2 inhibitor and approved for the treatment
of diabetes in 2017 (Roy and Ghosh, 2020; Bowles et al., 2014). More
2.4. Current trends in the field of pharmaceutical cocrystals recently, Mayzent, a cocrystal produced by Novartis, became the first
available treatment for secondary progressive multiple sclerosis, a form
The field of pharmaceutical cocrystals has experienced a steady of multiple sclerosis found in patients experiencing relapse (T-Cells). It is
growth over the past number of years with nine cocrystals being a 2:1 cocrystal of Siponimod, the API which is a sphingosine 1-phos­
approved by the FDA for commercial use, all of which are summarized in phate receptor modulator, and fumaric acid (Roy and Ghosh, 2020). It
Table 2. The first commercially available cocrystal, Beta-Chlor or chloral also exhibits polymorphism, but only the thermodynamically stable
betaine, achieved a US patent for commercial use in 1962. This was form I is sold commercially (EMA, Assessment report: Mayzent - Euro­
originally produced to mask the unpleasant taste of the API chloral pean Medicines Agency. Procedure No. EMEA/H/C/004712/0000,
hydrate, which is a sedative, but the increased thermal stability asso­ 2019).
ciated with this cocrystal was not realized until 2016, when it was first Table 2 illustrates the number of pharmaceutical cocrystals that have
described as a pharmaceutical cocrystal. This compound is formed by an been approved by the FDA, while Fig. 2 illustrates the number of pub­
exothermic reaction in water when heated to 60 ◦ C (Kavanagh et al., lications and patents per year in the field of pharmaceutical cocrystals. A
2019; O’Nolan et al., 2016; Petrow et al., 1962). Following the approval steady increase can be seen in the increase in publications and patents
of beta-chlor, Depakote was FDA approved in 1983 as a means of per year since 2004, with publications and patents displaying a 165 %
improving the solid form of the API and coformer by reducing their and 378 % increase respectively, between 2010 and 2020.
hygroscopicity. Depakote is a 3:1 cocrystal of sodium valproate and
valproic acid, and is used for the prevention of seizures in patients 3. Solid-state and particle engineering of cocrystals using
suffering with epilepsy (Petruševski et al., 2008). Another cocrystal atomization-based methods
whose structure was not uncovered until years later is the caffeine-citric
acid cocrystal. Approved as a complex in 1999 to treat infantile apnoea, Atomization is the process in which a solution (in the case of coc­
it was determined to be cocrystalline in 2007 (Kavanagh et al., 2019; rystals, one that contains an API and coformer) is broken into small
Karki et al., 2007). Lexapro (Escitalopram) is an antidepressant, formed droplets by high-speed fluids (e.g. CO2, N2, hot air) and then dried to
by escitalopram oxalate salt and oxalic acid. This cocrystal was found to produce a final powder. The selection on the nozzle type, orifice size and
be a more effective antidepressant than its enantiomer citalopram atomizing gas has a significant effect on the resulting particle sizes. One
(Harrison et al., 2007). Suglat was approved for use as a sodium glucose category of nozzles are pressure nozzles which allow for atomization by
transport 2 inhibitor in Japan in 2014 and is a 1:1 cocrystal of Ipragli­ pressurized solution flow through a small orifice which commonly
flozin and L-proline. It is used for the treatment of diabetes and similarly ranges from 10 − 1000 μm in diameter (Lefebvre and McDonell, 2017).
to the raw Ipragliflozin, the cocrystal with L-proline does not convert to The change from a high-pressure environment to a low pressure one,
a hydrate form under high stress conditions, and therefore has increased allows for the dispersion of a liquid stream into micro/nano-droplets and
long term stability (Roy and Ghosh, 2020). Entresto, approved for use in

Table 2
List of commercially available pharmaceutical cocrystals (Kavanagh et al., 2019; Roy and Ghosh, 2020).
Drug API Coformer Manufacturer Medical Dosage Year of Approval
Indication Administration

Beta- Chloral Hydrate Betaine Mead Johnson Sedation Oral (tablet) 1962 (O’Nolan et al., 2016)
Chlor®
Depakote® Sodium Valproate Valproic acid AbbVie Epilepsy Oral (tablet) 1983 (Petruševski et al., 2008)
Cafcit® Caffeine Citric acid Hikma Infantile Intravenous or Oral 1999 (Karki et al., 2007; Administration, F.a.D.,
Apnoea Caffeine citrate - New drug application (NDA 20-
793/S-001., 2000)
Lexapro® Escitalopram oxalic acid Forest Depression Oral (tablet) 2002 (Harrison et al., 2007; Masilamani and
oxalate salt Laboratories Ruppelt, 2003)
Abilify® Aripiprazole Fumaric Acid Otsuka Depression, Oral (tablet) 2006 (Devarakonda, 2009/0054455A1, 2007.)
Pharmaceutical Autism
Suglat® Ipragliflozin L-proline Astellas/ Diabetes Several solid and liquid 2014 (Imamura, et al., 2012)
Kotobuki formulations
Entresto® Sacubitril Valsartan Novartis Heart Failure Oral 2015 (Feng et al., 2012)
Odomzo® Sonidegib Phosphoric acid Novartis Basal Cell Oral 2015 (EMA, Odomzo, , 2015)
Monophosphate Carcinoma
Steglatro® Ertugliflozin L-pyroglutamic Pfizer/MSD Diabetes Oral 2017 (Bowles et al., 2014)
acid
Mayzent® Siponimod Fumaric acid Novartis Multiple Oral 2019 (FDA., 2019)
Sclerosis
Seglentis® Celecoxib Tramadol Kowa Acute Pain Oral 2021 (Elliott and Chan, 2021)
Pharmaceuticals

7
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

Table 3
List of commonly used supercritical fluids and their respective critical temper­
ature and pressure (Ruhan and Motonobu, 2009; Negoescu, 2017; Reid et al.,
1987).
Compound Critical Temperature (◦ C) Critical Pressure (MPa)

Carbon Dioxide (CO2) 31 7.4


Water (H2O) 374 22.1
Nitrogen (N2) − 147 3.4
Methanol (CH3OH) 239 8.1
Methane (CH4) − 83 4.6

and Motonobu, 2009; Negoescu, 2017). The low critical temperature of


CO2 is of particular interest as it allows for the processing of thermally
sensitive compounds, including (bio)pharmaceuticals. The solubility
behaviour of many substances in supercritical CO2 (scCO2) increases
with increasing pressure, which means that separation and purification
processes can be designed employing scCO2 as an extraction solvent.
Despite this tuneable solubility based on the pressure, the solubility of
most pharmaceutical drug molecules is low in scCO2, representing a
major challenge for methods that employ this scCO2 as a solvent. Su­
percritical CO2 is commonly used for extraction purposes, most
Fig. 2. Number of publications (red bars) and patents (green bars) per year commonly for the decaffeination of coffee in industry (Kim, 2008;
regarding pharmaceutical cocrystals since 2004, as determined by Scopus Sökmen et al., 2018) and for the extraction of essential oils and food
(adapted from(). products in academia (Sökmen et al., 2018; Khaw, 2017; Gallego et al.,
Kavanagh et al., 2019 2019). However, scCO2 can also be used for the production of phar­
maceutical drug particles. Supercritical CO2 has three unique roles
are primarily associated with conventional spray drying. Rotary atom­ which makes it widely versatile for the generation of pharmaceutical
izers on the other hand involve the injection of solution to the centre of a drug powders: as a solvent, antisolvent or co-antisolvent, and as an
nozzle chamber. As this chamber rotates at high speeds, the solution additive or spray enhancer. The selection on the scCO2 role to use is
flows to the periphery of this nozzle where it leaves through apertures primarily influenced by the solubility of the organic solvent and/or the
allowing for atomization. Air-assisted atomizers utilize high pressure gas solutes in scCO2. Each of these roles and the various processes associated
streams (usually air or nitrogen) to aid in solution dispersion resulting in with them will be discussed in greater detail in this review paper.
a fine spray of droplets. In these types of atomizers, the compressed gas
can come into contact with the solution after it has passed through the 3.1.1. Supercritical CO2 as a solvent
orifice (external gas–liquid mixing) or within the nozzle such that the The first role of scCO2 that will be considered is when the SCF acts as
gas and the solution leave the nozzle together as a mixed phase (internal a solvent to dissolve all, or part of the solutes involved in the final
gas–liquid mixing) (Lefebvre and McDonell, 2017). One such example formulation. This role can eliminate or greatly reduce the requirement
includes coaxial nozzles which are often used in supercritical fluid for organic solvents and thus reduce the number of steps required for
drying. This type of nozzle is composed of a nozzle pipe with a small drug processing. However, as a solvent, scCO2 faces some limitations as
diameter, within a pipe with a larger diameter. The solution flows the solubility of APIs vary in scCO2, many of which possess a low sol­
through the small diameter pipe while the supercritical medium flows ubility in scCO2. The solubility of APIs in scCO2 is influenced by its
through the larger pipe. In the absence of a mixer, this leads to external polarity with non-polar compounds exhibiting an increased solubility in
mixing but the inclusion of a mixer allows for internal mixing and the scCO2 compared to polar compounds. Due to the control over API/
atomization of the mixture of supercritical fluid and solution (Padrela coformer solubility through the use of co-solvents, precipitation times
et al., 2018). Other variations of atomizers have included the use of an can be controlled to a degree, and as such, it is possible to obtain various
electric charge as seen in electrospraying, ultrasonication or vibration to solid-state forms including pharmaceutical cocrystals. Additionally, due
improve aspects of the atomization step which will be later discussed in to the reduced solubility of APIs and coformers at lower pressures, it is
this review (Lefebvre and McDonell, 2017). Atomization provides a less likely that conversion will take place once precipitation has
large variety of options to tailor the drying process for the substances occurred (Padrela et al., 2018; Wang and Su, 2020).
being dried, such as droplets containing API molecules, by manipulating Rapid expansion of supercritical solutions (RESS)
the residence time and interaction between solvent, solutes, and atom­ The first method which employed scCO2 as a solvent was named as
izing fluid, such that processes such as cocrystallization is favoured. rapid expansion of supercritical solutions (RESS), which was first re­
There are multiple different methods which utilize the atomization ported in the 1980 s by Smith et al. (Matson et al., 1987; Matson et al.,
mechanism to produce pharmaceutical cocrystals and cocrystal solvates, 1986; Matson et al., 1987; Petersen et al., 1986; Smith, 1986). This
hydrates, polymorphs and cocrystals (and cocrystal polymorphs and method involves the dissolution of a solute(s) in scCO2 in a solubiliza­
solvates), which will be discussed below with respect to their unique tion vessel. The pressure and temperature of scCO2 can be adjusted to
advantages and challenges. achieve a saturated (or near-saturated) fluid solution. This solution
containing the dissolved API and coformer flows through a pre-
expansion zone and is then rapidly expanded through a capillary
3.1. Supercritical fluid-based methods nozzle into a collection chamber (expansion vessel), where CO2 is
removed via a vent. The rapid depressurization causes a high degree of
In this review, most of the supercritical fluid-based methods dis­ supersaturation which results in homogenous nucleation and causes fine
cussed use carbon dioxide (CO2) as the supercritical fluid, based on its solute particles (<1 μm) to precipitate with a narrow particle size dis­
mild critical parameters (Tc = 31.1 ◦ C and Pc = 7.38 MPa) (Padrela et al., tribution (Kumar et al., 2021). The time taken for precipitation to occur
2018). Table 3 lists the critical temperature (Tc) and pressure (Pc) of the was first estimated by Smith et al. to be in the range of microseconds and
most commonly used compounds for supercritical processing (Ruhan

8
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

this short time can favour the production of metastable forms, accord­ solution was atomized at a constant flow rate for 1 h through a 150 µm
ingly to the Ostwald’s rule of stages that states that the first form ob­ orifice nozzle, and the cocrystal particles were collected within a filter
tained is usually the least stable one (Matson et al., 1987; Petersen et al., medium. A slow solvent evaporation method was also used to produce
1986; Smith, 1986; Schmelzer and Abyzov, 2017; Matson et al., 1987). these same cocrystals, where the resulting powder was ground and
In order to accommodate APIs with poor solubility in scCO2, several passed through a 250 μm sieve to produce a cocrystal with a similar
adaptations of this method were created, which include a liquid solvent surface area for comparative dissolution studies with the unprocessed
being present in the expansion vessel (RESOLV/RESSAS) which aimed to ibuprofen (Mullers et al., 2015). The cocrystals produced by RESS and
reduce the amount of particle coagulation that was observed after solvent evaporation were equally pure with an identical molecular
passing through the nozzle, the use of a solid co-solvent (RESS-SC) constitution. Bother the physical mixture and cocrystallization of
which aided the solvation of APIs in the solubilization vessel, and the use Ibuprofen and nicotinamide proved to improve the solubility of
of a non-solvent (RESS-N) (Kumar et al., 2021; Xiang et al., 2019; Ibuprofen. However, dissolution rate of ibuprofen was only increased
Sodeifian et al., 2018). However, these variations have yet to be utilized after cocrystallization. The surface area of the cocrystals produced by
for cocrystal production. RESS were ten times larger than the unprocessed API, indicating a
Vemavarapu et al. investigated the coprecipitation of 12 different smaller particle size. Pure ibuprofen was also processed by RESS and
API-additive/coformer/dopant mixtures using the RESS process. Su­ displayed a surface area four times greater than the unprocessed
percritical CO2 was employed to demonstrate the ability of this process ibuprofen, again highlighting the ability of this method to greatly reduce
to produce various solid-state forms of APIs such as hydrates, cocrystals particle size. This size reduction led to an increase in dissolution rate
and amorphous forms. The RESS process utilized here tested both an compared to cocrystals prepared by solvent evaporation (Mullers et al.,
80:20 and 20:80 mixtures of the drug and additive/coformer/dopant in 2015).
the extraction vessel. Pressure and temperatures were varied for each of Overall, there are not a significant number of methods which employ
the mixtures to determine their impact on the produced products solid- scCO2 as a solvent coupled with atomization for the production of coc­
state (Vemavarapu et al., 2009). A reduction in crystallinity was rystals. This is primarily as a result of a significant majority of APIs and
observed in all cases. Vemavarapu et al. successfully cocrystallized sal­ coformers exhibiting poor solubility in scCO2. This type of scCO2 pro­
icylic acid with acetylsalicylic acid and benzoic acid. Cocrystallization of cessing (employing scCO2 as a solvent) can experience significant uptake
salicylic acid with the additive aspirin lead to a morphological change in the coming years as a large proportion of APIs are exhibiting greater
from long needles to what was described as “dense networks”. The in­ hydrophobicity. As discussed in an earlier section, currently scCO2
clusion of aspirin in the extraction vessel also allowed for a five-fold methods are extensively used to extract oils and lipid-like substances
improvement in salicylic acid yields in the final product due to co- based on their favourable solubility and this introduces a new solvent-
solute mediated solubility improvement, allowing for a more efficient free method to produce APIs. A variation of the RESS method is the
solubilization of salicylic acid in scCO2. It was also reported that a controlled expansion of supercritical solutions (CESS) method employed
powder with a low melting point was obtained for both the salicylic and patented by Nanoform, which have claimed to produce API particles
acid-acetylsalicylic acid cocrystal, and the salicylic acid-benzoic acid as small as 10 nm. Nanoform state that, unlike RESS, the CESS method
cocrystal, possibly due to the formation of a eutectic melt (Vemavarapu might provide significant control over the crystallization process
et al., 2009). The salicylic acid-acetylsalicylic acid cocrystal PXRD through controlled flow, mass transfer, pressure and particle collection
pattern demonstrated a sharp reduction in crystallinity, which the au­ in dry ice (Nanoform, What our technology can achieve., 2021; Pessi
thors proposed to be attributed to eutectic formation (Vemavarapu et al., 2016). Although the CESS method has not been reported for the
et al., 2009). This process also demonstrated the ability to control the production of cocrystals, it may be an attractive alternative to RESS for
formation of a hydrated or anhydrous cocrystal of theophylline and simultaneous cocrystallization and micronization.
caffeine. The formation of a cocrystal hydrate was observed at lower
temperature set conditions in the process which utilized a theophylline 3.1.2. Supercritical CO2 as an anti-solvent
and caffeine mixture. However, an increase in temperature or an in­ Using scCO2 as an anti-solvent is, by far, the most popular approach
crease in the caffeine composition (as high caffeine levels can compete with respect to drug processing using SCFs. This is primarily due to the
with water molecules during crystallization) in the starting material improved solubility and miscibility of organic solvents in scCO2
mixture prevented the formation of a hydrated crystalline form compared to APIs (Wu and Su, 2018; Said-Galiyev et al., 2004). Due to
(Vemavarapu et al., 2009). The RESS process has demonstrated relative the popularity of this role, there are multiple methods, many of which
control over hydrate formation where applicable, as well as the ability to include atomization, that have been developed for the production and
improve the resulting cocrystalline material solubility and yields processing of pharmaceuticals, including the production of cocrystals.
through the use of co-solutes (additives). However, a major drawback of The scCO2 antisolvent effect provides a unique mechanism to induce
this method and other supercritical solvent-based methods for cocrystal precipitation of APIs and coformers through extremely high supersatu­
production is the selective solubilization/extraction of the reactants. ration levels. Using scCO2 as an antisolvent can provide greater control
One example of this phenomenon is with theophylline-theobromine over the solid state (crystalline forms or amorphous) of the APIs, in
cocrystals. It was found that an increase in temperature and/or pres­ addition to other physical properties such as particle size and shape
sure led to an increase in the general amount of theophylline in the (Padrela et al., 2018; Franco and De Marco, 2021; Franco and De Marco,
cocrystals (Vemavarapu et al., 2009). The general selectivity of super­ 2020). Fig. 3 depicts a schematic of a typical supercritical antisolvent
critical solvents is an unfavourable feature for their use in pharmaceu­ method resulting in the production of cocrystals. The SCF solubilizes
tical crystallization, but Vemavarapu et al. demonstrated the ability to within the liquid solvent, reducing its solvation power and inducing
manipulate this downfall to a certain degree to promote desired crys­ supersaturation and thus, antisolvent nucleation. An atomization step
tallization, whilst improving yields and demonstrating a general control then allows the removal of the liquid solvent from the atomized feed
over the solid-state of the resulting cocrystalline materials. solution into the supercritical medium (Padrela et al., 2009). This sec­
The RESS method has also shown promise in its ability to simulta­ tion will investigate various scCO2 antisolvent methods which employ
neously (co)crystallize and micronize APIs due to the rapid depressur­ atomization as part of their mechanisms, and those that have demon­
ization (Sodeifian et al., 2018; Martello et al., 2019; Türk and Bolten, strated control over the solid-state and particle size of pharmaceutical
2016; Ghoreishi et al., 2016; Paisana et al., 2016). Müllers et. al. cocrystals. Gas antisolvent (GAS) crystallization is a popular method
attempted to produce cocrystals of ibuprofen and nicotinamide in a which has been utilized for the production of pharmaceutical cocrystals,
single RESS process for cocrystallization and micronization. The re­ but it does not involve an atomization-based micronization step and thus
actants were dissolved in supercritical CO2 over 24 h, which then the will not be discussed in detail in this section (Long, 2021; Pessoa, 2019).

9
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

rapid solvent evaporation. Similar to that which has been reported by


other studies (Zhao et al., 2018; Cuadra, 2020), the particle size of the
paracetamol-dipicolinic acid cocrystals produced by SAS (4.18 ± 0.054
µm) was smaller than that of the as-received paracetamol and the coc­
rystals produced by rapid solvent evaporation (45.33 ± 0.43 µm and
64.93 ± 0.095 µm, respectively). Reduced particle size generally leads
to an increase in cohesion leading to poor flowability, however the
opposite was observed in this study due to polyhedral particle
morphology of the cocrystal (compared to as-received paracetamol)
resulting in reduced particle contact points. Thus, these smaller coc­
rystals produced by SAS displayed improved flowability. However, the
flowability of the cocrystals produced by solvent evaporation was
slightly better than that of the SAS produced cocrystals due to their
larger particle size. The SAS process also produced cocrystalline pow­
ders with narrower size distribution than that of the solvent evaporation
process (Hiendrawan et al., 2016). Cocrystallization increased both
solubility and dissolution rate of paracetamol, but this increase was
higher in the SAS produced cocrystals due to their reduced particle sizes.
Due to the polyhedral morphology, the cocrystals displayed improved
compressibility and therefore tabletability, which was the primary aim
of that study (Hiendrawan et al., 2016).
Alternatively, Zhao et al produced a different cocrystal of paraceta­
mol with trimethylglycine by SAS, but unlike Hiendrawan et al. (Hien­
drawan et al., 2016), investigated the effect of changing SAS process
parameters and compared the resulting formulations with those of ball
milling. In preliminary experiments, it was found that the tensile
strength of the cocrystals was above the minimum value required for
Fig. 3. Schematic representation of a typical scCO2-antisolvent method that effective tabletting (2 MPa) and as such, the authors focused on
results in the generation of cocrystal particles. 1: Supercritical CO2 interacts and
improving the drug content in the final cocrystalline formulation. The
mixes with the liquid solvent that contains dissolved cocrystal reactants; 2:
optimized SAS process conditions increased tensile strength of the coc­
scCO2 is flushed out which contains the liquid solvent in which the reactants
rystal up to 11.13 MPa and displayed increased drug contents of
were initially dissolved in; 3: cocrystal particles precipitate out of the solution.
61.17%. Raw paracetamol and trimethylglycine showed the largest
particle sizes of approximately 400 μm and 90 μm, respectively, and the
The use of scCO2 as an additive or spray enhancer has also been used in
largest particle size distributions. The smallest particle size was obtained
combination with antisolvent methods. However, methods which only
from the optimized SAS sample of approximately 8 µm, whilst other
utilize the supercritical fluid as a spray enhancer will be discussed
particle sizes were not reported (Zhao et al., 2018). Samples produced
further in section 3.1.3.
by the SAS method displayed the optimum tabletability.
Supercritical antisolvent (SAS)
Neurohr et al. investigated the ability of the SAS process to produce a
One method that utilizes supercritical fluids as an antisolvent is the
pure cocrystalline powder of the naproxen-nicotinamide cocrystal. So­
supercritical antisolvent (SAS) method. The SAS method involves the
lution and scCO2 flow rates were varied, and in all cases the percentage
atomization of an organic feed solution, that the API and coformer has
of cocrystal produced ranged from 94 − 100%, demonstrating the ability
been dissolved in, into a pressurized precipitation chamber. The pre­
of the SAS method to produce highly pure formulations. The particle
cipitation chamber contains the supercritical CO2 at pre-determined
sizes of the initial runs ranged from 300 to 400 μm and remained
operating conditions and is introduced to the chamber through a sepa­
consistent through the variation of flow rates. An increase in solution
rate inlet to the solution (although a coaxial nozzle has been used in
concentration produced cocrystals with a similar particle size. The au­
combination with this in some cases (Padrela et al., 2009)) (Park, 2007;
thors of that study proposed that these similar particle sizes could be
Neurohr et al., 2016). The interaction of the atomized feed solution with
attributed to low supersaturation levels. As such, to promote higher
the scCO2 leads to volumetric expansion and subsequent precipitation of
levels of supersaturation, the weight-based flow ratio of CO2-to-solution
the solid cocrystalline particles (Park, 2007; Neurohr et al., 2016).
was increased from 11 to 36 g/g. This increase of supersaturation levels
As scCO2 is easily atomized and possesses density similar to liquids, it
resulted in reduced particle sizes to as low as 15 μm, but heterogeneity
is a very effective medium in an atomization process, aiding in the
was observed in the form of naproxen homocrystals. To combat this, the
breakup of a liquid stream into small droplets thus creating small par­
ratio of API to coformer was changed from 2:1 to 2:2 which produced a
ticles. For this reason, supercritical based methods such as the SAS
more pure formulation of cocrystal of approximately 99% (Neurohr
method have been effective in reducing particle sizes of pharmaceutical
et al., 2016). This paper demonstrated an ability of the SAS process to
substances (Montes et al., 2016; Yoon et al., 2016; Montes et al., 2016;
control and manipulate the solid state of the resulting dried formulation
Pan, 2020; Xu, 2020). The SAS method, in particular, involves the
through altering supersaturation levels and even the proportion of API
mixing of the supercritical fluid and the feed solution prior to atomi­
and coformer, allowing for the simultaneous reduction in particle size
zation, allowing for antisolvent nucleation, thereby increasing mass
and prevention of homocrystal production.
transfer which has a positive impact on particle size reduction (Vorobei
It is not often that a cocrystalline system can be produced by both a
and Parenago, 2021; Ha, 2020; Reverchon and De Marco, 2011). Table 4
supercritical solvent method as well as an antisolvent method due to
provides a list of all cocrystals produced using the SAS method to date,
their distinct modes of operation. Supercritical solvent-based methods
along with corresponding particle sizes.
require that both the API and coformer are approximately equally sol­
Hiendrawan et al. employed the SAS method in the production of a
uble in the supercritical fluid, whereas antisolvent methods require that
cocrystal of paracetamol and dipicolinic acid from a methanolic solution
the API and coformer are relatively insoluble in the supercritical fluid, a
(Hiendrawan et al., 2016). Equally pure cocrystal products were pro­
far more common instance. Cuadra et al. compared the production of
duced from the SAS process, compared to conventional methods such as
cocrystals of 5-flourouracil with various coformers (urea and thiourea)

10
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

Table 4
List of pharmaceutical cocrystals produced by various supercritical CO2-based atomization methods, along with the corresponding solvents and processing conditions.
Role of Method API Coformer Solvent Nozzle FSOL P (MPa) T (◦ C) Particle Reference
SCF Diameter (ml/ size
min) range

Solvent RESS Mefenamic acid Nicotinamide N/A 100 μm N/A 9.1 – 40 – 90 Not (Vaksler, 2021)
19.3 reported
Ibuprofen Nicotinamide N/A 150 μm N/A 30 50 Not (Mullers et al.,
reported 2015)
Salicylic acid Acetylsalicylic acid, N/A Not N/A 7.6 – 62 35–100 Not (Vemavarapu
Benzoic acid reported reported et al., 2009)
Acetylsalicylic acid Benzoic acid Not
reported
Tolbutamide Chlorpropamide Not
reported
Theophylline Theobromine, Caffeine Not
reported
Antisolvent SAS Paracetamol Nitroisophthalic acid Methanol 254 μm 1 10.0 40 4.66 μm (Tjandrawinata
et al., 2019)
Paracetamol Dipicolinic acid Methanol 254 μm 1 10.0 40 4.18 μm (Hiendrawan,
2016)
5-Flourouracil Urea Methanol 100 μm 1 8.0 40 10–60 (Cuadra, 2020)
μm
5-Flourouracil Thiourea Methanol 100 μm 1 10.0 40 5–65 μm (Cuadra, 2020)
Carbamazepine Saccharin Methanol, 100 μm 1 10.0 – 40.0 – 5 – 10 μm (Cuadra, 2018)
ethanol, 15.0 60.0
DCM, DMSO
Naproxen Nicotinamide Acetone 180 μm 2 – 13 10 37 300 – (Neurohr,
400 μm 2016)
Indomethacin Saccharin Ethanol, 80 and – 9.6 – 50 - (Padrela, 2009)
acetone, 200 μm 10.5
methanol,
THF, ethyl
acetate
Paracetamol Trimethylglycine DCM + 500 μm 1.2 10.0 45 - (Zhao, 2018)
methanol
Diflunisal Nicotinamide Acetone, 100 μm 1 10.0 – 35 – 40 Not (Cuadra, 2016)
ethanol 12.0 reported
AAS Indomethacin Saccharin Ethanol, 80 and – 6.0 – 50 – 70 0.5 – 0.7 (Padrela, 2009)
acetone, 200 μm 12.0 μm
methanol,
THF, ethyl
acetate
Theophylline, Saccharin Ethanol 100 μm – 8.0 – 8.2 50 Not (Padrela, xxxx)
Caffeine, reported
Sulfamethazine
Spray SEA Theophylline Urea, Saccharin, gentisic THF 100 μm – 4.0 – 40 – 70 0.5–0.7 (Padrela, 2014)
Enhancer acid, Salicylic acid, 10.0 μm
glutaric acid, sorbic
acid, 1-hydroxy-2-naph­
thoic acid, oxalic acid,
maleic acid,
nicotinamide
Theophylline Saccharin THF 100 μm – 8.0 50 Not (Tiago, 2013)
reported
Indomethacin, Saccharin Ethanol 100 μm – 7.8 – 8.3 50 0.3–10 (Padrela, 2010)
theophylline, µm
caffeine,
sulfamethazine,
aspirin
carbamazepine
Carbamazepine Saccharin Ethanol, Not Not Not Not Not (Padrela, 2019)
methanol reported reported reported reported reported

SCF (supercritical fluid), RESS (rapid expansion of supercritical solutions), SAS (supercritical antisolvent), AAS (atomization and antisolvent), SEA (supercritical
enhanced atomization), API (active pharmaceutical ingredient), DCM (dichloromethane), DMSO (dimethyl sulfoxide), THF (tetrahydrofuron), FSOL (solution flow
rate), P (pressure), T (temperature).

by the SAS and the CSS (cocrystallization with supercritical solvent) Methanol was employed as the solvent in the SAS method due to the
methods (Cuadra, 2020). The CSS method involves the solubilization of high solubility of the API and coformers in it. In the CSS method, as the
the solid components (API and coformer) in scCO2 in a high-pressure API and coformers have reduced solubility in scCO2, methanol was also
vessel, or the formation of a slurry my partially insoluble materials. used as a co-solvent due to its ability to increase the solubility of polar
This is followed by depressurization and subsequent induction of su­ compounds in scCO2 (Cuadra, 2020; Li, 2011; Bitencourt, 2019). The
persaturation which drives the precipitation of the cocrystals. Co- CSS method only produced cocrystals when methanol was included as a
solvents have been used in the past to aid in the solubilization of re­ co-solvent (to improve the amount of API and coformer that would be
actants in scCO2 (Padrela et al., 2009; Cuadra, 2020; Ribas, 2019). dissolved), which highlighted the importance of the solubility of the

11
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

starting materials in scCO2. The produced 5-flurouracil-urea cocrystal drying, as discussed further in section 4.2 (MacEachern et al., 2020).
particles exhibited plate-like shape, with a length of 10–60 µm, width of This method has been successful in the production of nano-sized lyso­
2–6 µm, and thickness of 10–20 nm (Cuadra, 2020). The lack of an at­ zyme particles, as well as several API cocrystals (see Table 4) which
omization step resulted in the production of much larger particles in the utilized saccharin as the coformer (Padrela et al., 2009; Rodrigues,
CSS method, comparatively to SAS, but the low solubility of reactants in 2009; Padrela). Padrela et al. produced indomethacin-saccharin coc­
scCO2 (even with the addition of an appropriate co-solvent) may have rystals using both the SAS and AAS method. Both methods utilized a
led to the formation of a mixed crystalline phase of cocrystals and coaxial nozzle with a 30 µl mixing volume for the induction of super­
homocrystals. The use of thiourea as the coformer in the SAS process led saturation and antisolvent nucleation. Within the AAS configuration, the
to the formation of a mixed phase of cocrystal and API homocrystals. SCF used can be altered between CO2 and N2, and other variations can be
These particles also displayed a plate-like shape with a particle length of made to the pressure, solvent used, and mass flow ratio of the SCF and
5–65 µm, width of 4–8 µm, and thickness of 0.5–3 µm in SAS. The CSS organic solution. Where nitrogen was utilized as the SCF, cocrystals
method also produced a mixture of cocrystal and API homocrystals for formed in the liquid feed solution and were further dried by the nitrogen
both urea and thiourea. Particle sizes for the CSS produced cocrystal gas. DSC and PXRD analysis confirmed the production of pure cocrystal
were not reported, but from the SEM images it is clear that the particle formulations, however enthalpy of fusion values showed minor varia­
lengths well-exceed 100 µm, which is in agreement with typical particle tions inferring the formation of polycrystals. The SAS-produced coc­
sizes reported when using the CSS method and surpassing those of the rystals displayed no significant changes in morphology when operating
reported SAS particle sizes (Padrela et al., 2018). Similar to the study conditions such as pressure or solvent were varied. These consistently
discussed above by Cuadra et al., the presence of homocrystals in the formed mixtures of block and needle-like particles with an approximate
precipitate is likely due to an incongruent saturating system, as urea and size of 5 μm. The AAS-produced cocrystals displayed a more spherical/
thiourea have higher solubilities in methanol than 5-flourouracil. granular morphology with a mean diameter of 500–700 nm, which was
Despite the addition of methanol, the production of homocrystals in also unaffected by changes in process conditions, solvent choice, or the
the CSS method is due to low solubilities of the solutes in scCO2. The co- SCF used. The larger size of SAS produced cocrystals may be attributed
solvent improved this as no cocrystals could be produced without it, but to the enhanced time in contact with the liquid organic solvent, which
it is likely that higher levels of solubility in the supercritical medium allowed for crystal growth, prior to solvent extraction by the super­
would be required for the CSS method (Cuadra, 2020). critical medium. The mixed particle morphology obtained in SAS was
A few cocrystals are known for exhibiting various polymorphic due to the crystallization mechanisms. These include the reduction of
forms. One example is the carbamazepine-saccharin (CBZ-SAC) cocrys­ the organic liquids solvent power by the dissolution of the SCF in the
tal, which exhibits two polymorphic forms, the stable form I and the solvent, and the removal of the solvent from the atomized droplets into
metastable form II. Cuadra et al. investigated the influence of operating the SCF. Whether these mechanisms occur simultaneously or not plays a
conditions such as pressure (10–15 MPa), temperature (40–60 ◦ C), sol­ large role in the final particle size and shape (Padrela et al., 2009).
vent choice (methanol, ethanol, dichloromethane (DCM) and dime­ Padrela et al. also successfully produced cocrystals using saccharin as a
thylsulfoxide (DMSO)), and API concentration (15–30 mg/ml) within an coformer with several APIs, including theophylline, caffeine, and sul­
SAS process on the polymorphic form of the 1:1 CBZ-SAC cocrystal. The famethazine using the AAS method and demonstrated the method’s
use of DMSO as a solvent produced no precipitate, as coformer solubility ability to produce less agglomerated particles than conventional
may have been enhanced in DMSO-CO2 system. PXRD confirmed the methods (Padrela).
presence of cocrystal form I when methanol was used, but a mixture of
forms I and II was observed when either ethanol or DCM was used. 3.1.3. Supercritical CO2 as a spray enhancer/additive
Operational conditions were then varied whilst using methanol as the Supercritical CO2 can also act as a spray enhancer or additive, such
solvent to improve yields. When a temperature of 40 ◦ C and pressure of that it may be a solute or a co-solvent. The unique properties of scCO2
10 MPa was used, a pure phase of form I was produced with yields of 65 make it an attractive candidate to feature as an additive or an atomizing
%, compared to 50 % under pressure conditions of 15 MPa. However, at fluid (spray enhancer). This SCF possesses the density similar to that of a
60 ◦ C a mixture of form I and II was produced, most probably due to liquid at mild pressures and temperatures compared to other SCFs (see
changes in the ternary phase diagram causing an incongruent saturating Table 3). Secondly, it is capable of dissolving in many substances such as
system. The higher temperature may have also allowed for a more rapid polymers, leading to a reduction in their melting temperature, such that
drying process which may have favoured the formation of metastable they can be used with thermally labile materials. The third and final
forms, according to Ostwald’s rule of stages (Schmelzer and Abyzov, feature, also observed in antisolvent-based atomization methods, is its
2017; Grossjohann, 2015; Walsh, 2018; Walsh, 2018; Tawfeek et al., ability to be easily atomized allowing the disruption of solution streams
2020). The authors pointed out that the reason methanol was the only into fine droplets, resulting in a greater level of control over the physi­
solvent to yield a pure sample of one of the cocrystal polymorphs is cochemical properties of pharmaceutical powders. This role of scCO2 is
because carbamazepine and saccharin have similar solubilities in commonly observed in combination with other previously mentioned
methanol, creating a congruently saturating system (Cuadra, 2018). This methods and can be regarded as complimentary to such methods. Ex­
has previously been noted to lead to the formation of homocrystals, but amples of supercritical spray enhancing methods include supercritical
in the case of carbamazepine and saccharin, it lead to the formation of a enhanced atomization (SEA), supercritical assisted atomization (SAA),
mixed phase of cocrystal forms I and II (Neurohr et al., 2016; Cuadra, and particle formation from gas saturated solutions (PGSS) (Campardelli
2018). et al., 2016; Peng, 2019; Ndayishimiye and Chun, 2018).
Atomization and antisolvent (AAS) Supercritical enhanced atomization (SEA)
The atomization and antisolvent (AAS) method has been considered The only method which utilizes the spray enhancing ability of SCFs
a derivative of the SAS method discussed above, but with the major for cocrystal production, is the supercritical enhanced atomization
difference being the precipitation chamber is held at atmospheric (SEA) method. This method is occasionally labelled as the supercritical-
pressure rather than high pressure. This method primarily utilizes a assisted spray drying method (SASD), but both SEA and SASD methods
coaxial nozzle with a small mixing chamber for the atomization of the appear to be somewhat identical in mechanism and design. The SEA
mixed scCO2/solution phase. This mixing allows for the induction of method is also the only method which has been used for cocrystal pro­
supersaturation and antisolvent nucleation. After the two feed streams duction which only utilizes the spray enhancer/additive role of SCFs.
mix for a small period of time, the homogenously mixed phase passes Unlike the AAS method, the mixing volume and time are low which does
through the nozzle, producing the atomized feed stream which, unlike not allow for any antisolvent crystallization, only antisolvent nucleation
SAS, is further dried by thermal means similar to conventional spray in some cases (Jung and Perrut, 2001; Long, 2019). The SCF low

12
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

viscosity and high density allows them to be effective in breaking up 3.2. Conventional spray drying
liquid streams to allow for effective micronization and simultaneous
drying. This method has shown promise in vaccine production, cocrys­ Spray drying refers to the atomization of a feed solution followed by
tallization, and lipid dispersion of cocrystals (Padrela et al., 2010; Tiago, thermal drying by a heated stream of gas. Nitrogen is the most
2013; Rodrigues, 2012). commonly used atomizing gas, particularly when using organic solvents,
The SEA method was first introduced in a 2010 publication for the but heated air can be used when the proportion of organic solvent in the
screening of pharmaceutical cocrystals (Padrela et al., 2010). Saccharin atomized solution is lower than 20 % v/v. After atomization, the feed
was utilized as the coformer in the screening of cocrystal systems which solution meets the hot gas in either a co-current configuration (same
included the APIs indomethacin, theophylline, caffeine, sulfamethazine, flow direction) which allows for less particle agglomeration, or a
aspirin, and carbamazepine in methanol solutions. With regards to the counter-current (feed and gas flow in opposite directions) configuration
theophylline-saccharin screening tests, two molar ratios were investi­ which allows for a reduction in energy input for evaporation (Patel et al.,
gated, 1:1 and 1:2. Analysis by PXRD and DSC highlighted the presence 2009; Piatkowski and Zbicinski, 2006). The geometry of the drying
of the 1:1 cocrystal, as well as a new cocrystal with reactant ratios of 1:2. chamber, where the solvent is removed from the precipitated API, can
All other APIs tested produced pure cocrystalline formulations. Particle have significant effect on the final solid-state form, particle size and
sizes of cocrystals produced by SEA were compared to conventional morphology. Vertical drying chambers are more commonly used than
methods such as grinding. Cocrystal system produced by SEA were horizontal chambers, but atomizers such as the rotary atomizer are more
found to have smaller median particle sizes compared to those produced effective in horizontal drying chambers due to the wide spray formed
by grinding. Specifically, theophylline-saccharin cocrystals produced by (Patel et al., 2009; Mujumdar et al., 2010). A drying chamber may also
SEA had a mean particle size of 500 nm, whereas grinding produced be considered a single stage or two-stage drying chamber. Two-stage
cocrystals with a mean particle size of approximately 700 nm, demon­ drying chambers allow the gas and feed solution to mix prior to atom­
strating the impact of atomization as well as scCO2 on reducing cocrystal ization into the drying chambers, while a second feed of hot gas further
particle sizes (Padrela et al., 2010). dries the atomized feed. This drying process allows for lower tempera­
A publication which further built upon the above mentioned article ture and is more suited for thermolabile products. Single stage drying
(Padrela et al., 2010) sought to produce theophylline cocrystals with does not include this mixing step (Patel et al., 2009; Mujumdar et al.,
various coformers (see Table 4) to achieve a cocrystal with appropriate 2010). Smaller droplets can produce smaller particles but if the drying
dissolution kinetics for medical use (Padrela et al., 2014). This paper process has not been altered accordingly, the particles can become
determined that the theophylline cocrystals solubilities were dependent damaged by prolonged exposure to the hot gas. Similarly, large droplets
on the coformer solubility, as the coformers with aqueous solubilities may produce larger dried particles but steps must be made to ensure that
lower than that of the API (gentisic acid, hydroxy-2-naphthoic acid, those particles are fully dried before collection (Patel et al., 2009;
saccharin, salicylic acid), generated cocrystals with aqueous solubilities Cotabarren, 2018). Bertin et al. noted that air flow rate and feed solution
lower than that of the API (Padrela et al., 2014). In line with this, concentration had large impact on the droplet size and thus, the size of
coformers with high aqueous solubilities (urea, glutaric acid, maleic the resulting ciprofloxacin hydrochloride particles (Cotabarren, 2018).
acid, nicotinamide, oxalic acid, sorbic acid) generated cocrystals with Conventional spray drying has been employed for cocrystal pro­
higher solubilities than that of the API, theophylline. Particle sizes and duction in recent years. Weng et al. investigated the effects of various
morphology were measured for the cocrystal systems to determine the parameters within a spray drying configuration on resulting particle size
effect of process variables such as solution concentration, temperature, of cocrystals of itraconazole and suberic acid. The operating conditions
and pressure. All cocrystals produced by the SEA method presented an that were examined in that study included gas flow and solution con­
average size between 500 and 700 nm indicating that coformer choice centration (Weng, 2019). Gas flow has been reported to impact the at­
had little impact on particle size. Coformer choice did show an effect on omization rate which may generate smaller particles, while solution
cocrystal morphology. Cocrystals with gentisic acid and saccharin pro­ concentration may affect the quantity of solute (cocrystal) inside each
duced granular/plate-like particles, whilst nicotinamide and sorbic acid droplet. (Karashima, 2017). Lower solution concentrations may result in
coformers produced cocrystals with a needle-like shape. Variation in smaller particle sizes due to less solute being present in each droplet,
solution concentration or pressure had little impact on the resulting while a larger atomizing gas flow rate would allow for an increased rate
particle sizes. However, temperature did show an effect on particle size, of atomization leading to smaller droplets (Karashima, 2017; Littringer,
as temperatures below the solvent boiling point yielded particles of 2013; Arzi and Sosnik, 2018; Baumann et al., 2021; Lu, 2022). Addi­
500–600 nm, whilst 800–900 nm particles were generated at tempera­ tionally, the solution concentration may influence the supersaturation at
tures above the solvent boiling point (Padrela et al., 2014). the point of atomization and can play a role in controlling the solid-state
Subsequently, Padrela et al. investigated the simultaneous cocrys­ of the API atomized, in that low supersaturation levels might lead to
tallization and lipid dispersion of the theophylline-saccharin cocrystal more stable forms while high supersaturation levels might lead to
using the SEA process, to improve the stability of the API. To produce the metastable or amorphous forms (Grossjohann, 2015; Walsh, 2018;
lipid loaded cocrystals, solutions of the cocrystal reactants and hydro­ Walsh, 2018; Tawfeek et al., 2020). In this study, Weng et al., gas flow
genated palm oil were atomized with scCO2 into the heated precipita­ rate were set to either 601 or 742 Nl/h, while solution concentration was
tion chamber with a filter for particle collection. The PXRD varied between 1.14, 3.41 and 5.68 mmol/L. In general, reduced solute
diffractograms confirmed the cocrystals purity. Cocrystallization concentration and increased gas flow rate led to a decrease in particle
significantly improved the stability of the API under 92% humidity size, but the smallest particle size of 2.56 μm was observed using solu­
conditions from several days to over 6 months. This was observed with tion concentrations of 3.41 mmol/L and a gas flow rate of 742 Nl/h.
and without lipid dispersion. Lipid dispersion also slowed the release of Whilst larger gas flows produced smaller particles, larger particle sizes
the API, preventing the formation of its hydrated form (Tiago, 2013). of approximately 4 μm were produced in runs which utilized solution
Although the SEA method has not yet demonstrated a great ability to concentrations of 1.14 mmol/L (Weng, 2019). This was attributed to
control the solid-state of cocrystalline systems apart from an ability to particle agglomeration, which is often seen in smaller sized crystals due
produce a pure cocrystalline formulation, it has shown an ability to to their electrostatic nature (Weng, 2019; Karashima, 2017).
produce polymorphs of cocrystals which have not been produced by Several publications have demonstrated the ability of a spray drying
other methods. This method has also been effective in reducing cocrystal configuration to produce a(n) (co)amorphous pharmaceutical product
particle sizes and particle size distributions in comparison to conven­ (Walsh, 2018; Lenz, 2017; Wicaksono, 2021; Fung and Suryanarayanan,
tional methods, such as grinding. 2019; Matsuda, 1984; Lyu et al., 2017; Jog and Burgess, 2018). Tawfeek
et al. demonstrated the ability to control the solid state of the

13
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

indomethacin-nicotinamide system to produce wither the cocrystal or a cocrystal suspension, the moisture content coupled with a high aspirator
fully coamorphous product (Tawfeek et al., 2020). That study aimed to rate could attribute the high level of crystallinity to retained cocrystal.
produce fully coamorphous products by spray drying homogenous sus­ Apart from this run, the next highest level of crystallinity achieved was
pensions of the indomethacin-nicotinamide cocrystal, whilst varying 7.86 % (Tawfeek et al., 2020). This run utilized low concentration and
solution concentration, pump rate, inlet temperature and aspirator flow low temperature allowing for a slower drying process, which is in
rate to influence the crystallinity of the produced product. Inlet tem­ agreement with the theory above in that these conditions induce lower
perature has been shown to influence the crystallinity of a final phar­ levels of supersaturation, keeping the product more crystalline (Tawfeek
maceutical product. Higher temperatures lead to an increase in droplet et al., 2020). The manipulation of processing conditions to influence
solvent evaporation which, as mentioned earlier, leads to higher su­ supersaturation levels is an effective method to control product crys­
persaturation levels which would typically promote the formation of tallinity, particularly during spray drying due to the already rapid
metastable and coamorphous/amorphous products (Lyu et al., 2017; evaporation process.
Jog and Burgess, 2018; Jensen, 2016). Assumingly, an increase in so­ Excipients have also shown an effect on cocrystal solid state, as
lution concentration leads to further high supersaturation levels result­ demonstrated by Walsh et al. The difference in Hansen Solubility Pa­
ing in a coamorphous product, whilst a reduced solution flow rate would rameters (HSP) has been employed as a general indication as to the solid
also indicate less solute in the chamber causing a more rapid evapora­ state of the spray dried formulation (Walsh, 2018). Low ΔHSP values
tion of solvent droplets. Tawfeek et al. reported the lowest degree of between API and coformer is a strong indication of cocrystal production,
crystallinity of 0.77 % in the final product occurred using the highest as the two components are easily miscible, which is favourable for
solution concentration values of 10 mg/ml along with the lowest solu­ cocrystal formation (Salem et al., 2019; Walsh, 2018; Mohammad et al.,
tion flow rate and highest inlet temperatures of 15 % and 150 ◦ C, 2011). However, it has been reported that a difference in HSP values
respectively, again due to rapid solvent evaporation. All runs had a large between cocrystal and the excipient greater than 9.6 MPa0.5 (ΔHSP >
reduction in crystallinity (<8 % crystallinity remaining) except one run 9.6 MPa0.5), is a strong indication of cocrystal formation, whilst values
which displayed 43.54 % crystallinity. This is believed to have been below 9.6 MPa0.5 indicate an unlikelihood towards cocrystal production.
caused by a relatively high moisture content, coupled with the high These low ΔHSP values indicate a high miscibility between cocrystal
aspirator rate. As the feed solution employed in this study was a and excipient which, in turn, blocks interactions between the API and

Table 5
List of pharmaceutical cocrystals and coamorphous pharmaceutical formulations produced by conventional spray drying, along with the corresponding solvents and
processing conditions (years 2015–2021).
API Coformer Solvent Nozzle FSOL FGAS Aspiration Outlet Particle size Reference
Diameter (ml/ (unit/ rate (m3/h) (◦ C) range (μm)
(mm) min) h)

Itraconazole Suberic acid Ethanol + – 1.5 601 – 35 50 2.56 ± 2.27 (Weng, 2019)
chloroform 742 NL
Sulfadimidine 4-aminosalicylic Ethanol – 9 – 10 667 NL 35 50 – 57 1 – 10 (Walsh, 2018)
acid
Indomethacin Nicotinamide Ethanol + water – – – – – - (Tawfeek et al., 2020)
Indomethacin Arginine Acetone, Water 0.7 6 40 m3 – 50 0.169 (Lenz, 2017)
Ketoconazole Oxalic acid Methanol – 3.5 – – 35 - (Fung and
Tartaric acid Suryanarayanan,
Citric acid 2019)
Succinic acid
Atorvastatin Isonicotinamide Methanol – 5 600 L – 55 - (Wicaksono, 2021)
calcium Maleic acid
Indomethacin Arginine Acetone + water – 6 667 L/h – 50 - (Jensen, 2016)
Histidine
Lysine
Sulfadimidine 4-aminosalicylic Ethanol – 9 – 10 473 NL 35 50 – 57 - (Grossjohann, 2015)
acid
Niclosamide Urea IPA 0.7 3–6 246 L 30 40 – 88 2 – 3.4 (MacEachern, 2021)
Niclosamide Nicotinamide Ethanol 0.7 1 – – 40 – 45 1–5 (Ray, 2020)
Cilostazol 4 HBA Acetone + – – 1 kg – – 53 – 91 (Urano, 2020)
2,4-di-HBA methanol 63 – 85
2,5-di-HBA -
Carbamazepine Saccharin Ethanol + 0.7 – – – – - (Padrela, 2019)
methanol
Caffeine Dapsone Acetone, – 10 ml/ 670 L – – ~5.1 – 5.4 (do Amaral, L.H., ,
ethanol, ethyl min 2018)
acetate
Ketoconazole Oxalic acid Methanol – 3.5 – – 35 - (Fung, 2018)
Tartaric acid
Citric acid
Succinic acid
Sulfadimidine 4-aminosalicylic Ethanol – 9 – 10 473 NL 35 50 – 57 5.7 ± 0.7 (Serrano, 2016)
acid
Sulfadimidine 4-aminosalicylic Ethanol 0.7 9–10 473 NL – 50 – 57 5.7 ± 0.7 (Serrano, 2016)
acid
Sodium Lactose Water – 3 40 m3 – 63 – 65 (Sovago, 2016)
naproxen
Theophylline Saccharin Methanol – 5 357 L – 50 – 55 <5 (Alhalaweh, 2013)
Urea Methanol – 5 357 L – 50 – 55 <5
Nicotinamide Water – 5 357 L – 50 – 55 <5

API (active pharmaceutical ingredient), HBA (hydroxybenzoic acid), FSOL (solution flow rate), FGAS (atomizing gas flow rate).

14
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

coformer (Walsh, 2018). This was shown to be the case regarding spray due to their rapid solvent evaporation ability. However, the added
drying to produce the sulfadimidine – 4-aminosalycilic acid cocrystals. combined vibrational energy of the formed Taylor cone during the
Any excipient used, regardless of whether they were amorphous or electrospraying method increases this nucleation rate further (Patil
crystalline by nature, produced cocrystals when ΔHSP values were et al., 2017). Another primary advantage of this method over other
greater than 9.6 MPa0.5. A mixture of cocrystal and reactants was found atomization-based methods discussed here is the higher yields achieved,
when using glycine and polyvinyl alcohol (PVA) as excipients which as particles are attracted to and collected on an inversely charged plate
displayed ΔHSP values of 6.6 and 4.9 MPa0.5, respectively. Hydrox­ (Patil et al., 2016; Patil et al., 2017; Rodrigues, 2018).
ypropyl methylcellulose (HPMC), polyvinylpyrrolidone (PVP) and sol­ This method has shown an ability to produce various polymorphic
uplus (which produced ΔHSP values of 1.9, 4.4 and 3.9 MPa0.5 crystal forms (Padrela et al., 2019; Patil et al., 2016; Patil et al., 2017;
respectively) resulted in fully amorphous products. However, this is not Patil et al., 2017; Al-Ani, 2020). Sharvil Patil (Patil et al., 2016; Patil
a golden rule. Chitosan was one tested excipient which proved to be an et al., 2017; Patil, 2018; Patil et al., 2018) first produced cocrystals by
exception, as it led to the production of cocrystal mixed with reactants this method in 2016, and demonstrated the ability of the method to
while exhibiting a ΔHSP value 11.2 MPa0.5. This occurred due to in­ produce various polymorphic cocrystal forms (Patil et al., 2016). Patil
teractions between this alkaline polymer and the acidic coformer 4-ami­ et al. aimed at the production of caffeine and maleic acid cocrystals.
nosalycilic acid (Walsh, 2018). Three polymorphic forms of this cocrystal exist, forms I and II of the 1:1
Spray drying has also demonstrated an ability to produce different cocrystal, and form I of the 2:1 cocrystal. Solvents chosen for this system
polymorphic forms of pharmaceutical cocrystals. A summary of pro­ were acetone, methanol, ethyl acetate and distilled water. It is less en­
duced cocrystals is listed in Table 5. The rapid drying process observed ergy efficient to use water for such a process, as it has a relatively large
during spray drying can allow for the production of less stable or surface tension, requiring a high voltage of 20 kV for effective atomi­
metastable polymorphs (Grossjohann, 2015; Walsh, 2018; Lu, 2022). zation. This high voltage was used for all other solvent systems. Despite
Grossjohann et al. demonstrated the ability of spray drying to produce a the reactant ratio in the feed solution (1:1 or 2:1), both forms I and II of
new polymorphic form of the sulfadimidine – 4-aminosalycilic acid the 1:1 cocrystal formed when using solvents ethyl acetate and acetone.
cocrystal (form II) compared to the previously reported form I produced Unlike methanol and water, ethyl acetate and acetone are aprotic sol­
by LAG (liquid-assisted grinding) (Grossjohann, 2015). Over the vents lacking acidic hydrogen groups and thus, produced 1:1 cocrystals
extended period of time assessed, both polymorphs displayed similar as they could not act as hydrogen bond donors. When using water as the
stability retention, but it would be hypothesized that the rapidly spray solvent, the 2:1 cocrystal was produced. This may be attributed to the
dried polymorph may be less stable over a longer period of time. stronger attractive forces of maleic acid with water, allowing for a more
predominant caffeine precipitation, resulting in the formation of 2:1
caffeine-maleic acid cocrystals. However, the volatility of the other
3.3. Electrospraying/electrohydrodynamic atomization solvents compared with water may also play a role in this. The highly
volatile nature of methanol, acetone and ethyl acetate compared to
Electrospraying (also call electrohydrodynamic atomization) is a water may have influenced the generation of the 1:1 cocrystals as crystal
relatively new method which has been employed for the production of growth occurs much more rapidly in these volatile solvents (Patil et al.,
cocrystals of pharmaceutical compounds including energetic cocrystals 2016). Solvent choice with this cocrystal has been noted in the past
for use as explosives (Patil et al., 2016; Huang, 2018; Patil et al., 2017; (Leyssens, 2012). Padrela et al. also highlighted the innate ability of the
Patil, 2018; Patil et al., 2018; Emami, 2018; Emami, 2018; Solaimalai, electrospraying method to generate metastable cocrystal forms, unlike
2019). A summary of all pharmaceutical cocrystals produced by this other methods. Carbamazepine-saccharin cocrystals were produced by
method is presented in Table 6. The mechanistic details of the electro­ various methods from either ethanol or methanol solutions. Solvent
spraying method involves electric forces coupled with the atomization of evaporation from methanol produced the metastable form II of
a feed solution of API and coformer into charged droplets. This form of carbamazepine-saccharin cocrystal, whilst the stable form I was pro­
atomization forms a Taylor cone with high vibrational energy at the tip duced from ethanol. This may be attributed to the lower boiling point of
of the atomizer needle. This then transforms into a cone jet which methanol, allowing for a more rapid evaporation. Spray drying pro­
further propagates into a very fine mist of charged droplets. Optimiza­ duced a mix of forms I and II from both solvents but the percentage of
tion of process parameters such as potential difference, solution flow form II present increased with the use of methanol, again attributed to a
rate and working distance (distance between needle and grounded lower boiling point and therefore accelerated evaporation (Padrela
collection plate) is required to ensure this mist of charged droplets forms et al., 2019). This was also aided by the larger droplets generally
(Emami, 2018). Atomization methods afford for a high nucleation rate

Table 6
List of pharmaceutical cocrystals reported to date produced by the electrospraying/ electrohydrodynamic atomization method, with the corresponding solvents and
operating conditions.
API Coformer Solvent Voltage Working FSOL T Particle size Reference
(kV) distance (cm) (ml/h) (◦ C) (μm)

Caffeine Maleic acid Methanol, acetone, ethyl 20 25 2 – - (Patil et al.,


acetate, water 2016)
Carbamazepine, Nicotinamide Methanol, ethanol 20 25 2 40 - (Patil et al.,
Itraconazole 2017)
Forskolin Ascorbic acid, Ethanol 15 20 2 40 1–2.5 (Patil, 2018)
nicotinamide, caffeine
Quercetin Caffeine, Nicotinamide Methanol 15 25 2 40 - (Patil et al.,
2018)
Indomethacin, Saccharin, Nicotinamide, Methanol, Ethanol, 20 20 2.4 – 0.219–1.000 (Emami, 2018)
Naproxen Isonicotinamide Acetone, Ethyl Acetate
Plumbagin Nicotinamide Methanol 15 25 1.5 25 - (Solaimalai,
2019)
Carbamazepine Saccharin Ethanol + methanol – 14 – – - (Padrela,
2019)

API (active pharmaceutical ingredient), FSOL (solution flow rate), T (temperature).

15
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

observed in conventional spray drying allowing for polymorphic trans­ 2018).


formation (Padrela et al., 2019; Halliwell, 2017). Supercritical nano The main advantage that the electrospraying and other atomization-
spray drying produced form I cocrystals only, whilst electrospraying based methods display over other cocrystallization methods, which do
produced from II only (Padrela et al., 2019). This again shows the ability not include an atomization step, is their ability to atomize the feed so­
of electrospraying over other established methods, to consistently pro­ lution to control the particle size of cocrystals of both congruent and
duce pure metastable polymorphic forms of pharmaceutical cocrystals. incongruent systems, highlighting the power of this method, negating
Emami et al. first investigated the production of nano-sized cocrys­ the need to construct ternary phase diagrams (Emami, 2018). Fig. 4
tals via the electrospraying method. The specific cocrystal systems below summarizes the strengths, weaknesses, opportunities and threats
looked at include indomethacin with saccharin or nicotinamide, and of the four main atomization-based methods for cocrystal production.
naproxen with nicotinamide or isonicotinamide. The electrospraying
method was successful in producing all of these cocrystals, except
indomethacin-nicotinamide which was found to be coamorphous 3.4. Other methods
(Emami, 2018). This may be attributed to the rapid evaporation
observed in this method, coupled with the propensity for indomethacin 3.4.1. Spray flash evaporation
to form amorphous/metastable forms after electrospraying and other Spray flash evaporation is another drying/crystallization method
rapid evaporation methods such as spray drying (Padrela et al., 2019; which has been sparsely investigated within the pharmaceutical sector
Nyström et al., 2011; Karmwar, 2011). For all successfully produced regarding cocrystal production but has been looked at extensively with
cocrystals by electrospraying, pure cocrystal powders were formed its use in the production of other substances such as explosive materials
regardless of whether the saturation conditions were congruent or (Huang, 2018; Li, 2015; Spitzer, 2014). In recent years, spray flash
incongruent (Emami, 2018). Whether a solution is congruent or incon­ evaporation has garnered attention for its potential for water desalina­
gruent is determined by the solubility of the cocrystal components at a tion (Coty, 2020; Chen, 2018). Spray flash evaporation begins with the
set temperature in a particular solvent, and the concentration of the heating and pressurization (using nitrogen) of the solution containing
reactants at the eutectic point. A congruently formed cocrystal is ther­ the desired material. The solution is then sprayed through a heated
modynamically stable and can be formed readily, whereas an incon­ atomizer into the drying chamber, which is held at a pressure lower than
gruently saturating system produces a less stable solid form due to solid- that saturation pressure of the solution. Flash evaporation occurs,
state transformations (Alhalaweh and Velaga, 2010). Particle size was allowing the removal of the solvent as a vapor, as observed in Fig. 5
also analysed in this study and the smallest cocrystal particles were that (Coty, 2020; Chen, 2018). This method has shown potential for the
of indomethacin-saccharin when processed with acetone. This may be in production of nano-sized APIs such as crystals, cocrystals, and nano-
part due to acetone possessing the lowest boiling point of all solvents sized core shell structures (Coty, 2020). Spitzer et al. produced nano-
used in this study, allowing for rapid evaporation (Emami, 2018). Unlike sized cocrystals for the first time using the spray flash evaporation
the needle-like micron-sized particles obtained from solvent evapora­ method. 2:1 cocrystals of caffeine and oxalic acid, were produced along
tion, the indomethacin-saccharin cocrystal produced by electrospraying with 1:1 cocrystals of caffeine and glutaric acid. Acetone solutions of
displayed a morphology dependent on the concentration of the initial cocrystal components were prepared in a low boiling solvent as
feed solution. At low concentrations (7.5 mg/ml), porous block-like required. The solution was atomized at a pressure of 4.0 – 6.0 MPa
particles with a mean diameter of 219 nm were produced. Conversely through a heated hollow-cone atomizer into a 0.5 MPa atomization
at higher concentrations (15 mg/ml), µm-sized needle-like particles chamber. The sharp and sudden pressure drop produced a thermody­
were produced along with 400 nm block-like particles. Two naproxen namically unstable solution which regained its stability through the
cocrystals were also produced with nicotinamide and isonicotinamide conversion of excess energy into latent energy, causing accelerated
coformers, both of which displayed a plate/flake-like shape (Emami, evaporation. A coupled temperature drop prevented the growth of the
produced cocrystals. PXRD confirmed cocrystal production and all mean

Fig. 4. SWOT (Strengths, Weaknesses, Opportunities, Threats) analysis of the four primary atomization-based techniques utilized for cocrystal production (Karimi-
Jafari et al., 2018; Padrela, 2018; Emami, 2018; Emami, 2018).

16
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

Fig. 5. Illustration describing the mechanisms of solvent removal in spray-freeze drying, spray congealing and spray flash evaporation.

particle sizes were found to be in the nano-size range, with caffeine- 3.4.3. Spray-freeze drying
glutaric acid cocrystals displaying a mean particle size of 111 nm Spray-freeze drying is a combination of the atomization ability of
(Spitzer, 2014). Unfortunately, no particle size measurements were re­ spray drying coupled with the sublimination drying process seen in
ported for the caffeine-oxalic acid cocrystal (Ghosh, 2020). lyophilization. This is a method which is capable of producing porous
particles at a reduced temperature, as compared to spray drying and
3.4.2. Spray congealing spray congealing, making it suitable for thermolabile molecules and
Spray congealing, also described as spray cooling or spray chilling, is biomolecules (Tanaka, 2020; Eddleston, 2013). A general process flow
a technique which combines aspects of spray drying technology with hot for spray-freeze drying begins with the atomization of a feed solution
melt extrusion and can be tailored to be both a scalable and ‘green’ directly into a source held at a reduced temperature. Liquid nitrogen is
method for pharmaceutical particle engineering (Albertini, 2008). With commonly utilized for this. In the case of cocrystal production, this
the purpose of cocrystallization, this process begins by homogenously produces frozen droplets of API, coformer and solvent. A freeze dryer is
mixing the API and coformer in a heated vessel to produce a molten feed then utilized to sublime the solvent by reducing the pressure below the
solution. This solution enters a jacketed feed line and is atomized solvents’ triple points and by applying a gentle heat. This typically
through a heated nozzle (Duarte, 2016). Typically rotary atomizers are generates a dried, porous and partially amorphous product (Tanaka,
utilized in spray congealing but these generate a wide spray and, as a 2020; Eddleston, 2013). To the best of our knowledge, Tanaka et al. is
result, a wide drying/evaporation chamber is required (Albertini, 2008). the only researcher who has published work on cocrystallization by
A cooled gas, typically nitrogen, is used to solidify the atomized molten spray-freeze drying. The aim of that study was to produce 2:1 cocrystals
feed and, similarly to spray drying, can be carried out in a co-current or of theophylline and oxalic acid for use as inhalable therapeutics, with an
counter-current configuration. A cyclone, filter or other capture mech­ increased stability within humid environments during storage or in the
anisms can be utilized for the separation of the solid particles from the lungs themselves. Water was utilized as the solvent and a 2-fluid nozzle
off gas (Duarte, 2016). The first and only reported paper on the pro­ carried out atomization. Spray dried cocrystals were utilized as a means
duction of cocrystals by spray congealing was carried out by Duarte et al. of comparison and the median particle size determined by SEM images
as a solvent free approach. These authors investigated the production of was not statistically different between the two processes (7.00–7.20
two caffeine cocrystals, with glutaric acid and salicylic acid as coform­ µm). However, due to the more porous structure of the spray freeze dried
ers, as well as the carbamazepine-nicotinamide cocrystal. Melting samples, mass median aerodynamic diameter (MMAD) was smaller in
occurred within a mixed, heated vessel until the total 30 g of API and spray-freeze dried samples than in spray dried samples, 3.03 μm and
coformer melted. A jacketed two-fluid nozzle was utilized for the at­ 3.46 μm respectively. However, all powders produced by spray-freeze
omization of the molten feed mixture and nitrogen was utilized for drying had a lower level of crystallinity than that of spray dried pow­
droplet solidification in a co-current mode. Inlet and outlet temperatures ders due to hygroscopicity, but cocrystallization itself overcomes this
varied depending on the nature of the reactants within the feed melt. negative effect by reducing moisture absorption. This allows for greater
The caffeine-salicylic acid and carbamazepine-nicotinamide cocrystals control of surface hygroscopicity and good pulmonary delivery. The
were utilized to determine the feasibility of the process to produce cocrystals displayed increased stability under high humidity conditions
pharmaceutical cocrystals, and their presence was confirmed by PXRD in comparison to the API. The freeze-drying stage of the process allowed
and DSC. Mean particle sizes of the caffeine and carbamazepine coc­ for the formation of hydrogen bonds between theophylline and oxalic
rystals were 13.59 μm and 31.56 μm, respectively, and were composed acid which prevented the formation of a theophylline monohydrate
of agglomerates of needle-like cocrystal particles. Inlet temperature form. Oxalic acid is preferred to water molecules as the molecular dis­
influenced the drying capacity of the method to produce fully dried tance as well as the energies for reaction are lower during cocrystal
caffeine-glutaric acid cocrystals, whilst atmospheric pressure influenced formation rather than hydration, thus the cocrystal is hydroscopically
the final particle size. It was noted that larger molten droplets (and thus more stable (Tanaka, 2020). Although spray drying is a more robust
larger final particles) led to the production of higher amounts of reactant process with the ability to utilize various solvents and produce particles
impurities in the final formulations, most likely due to an inefficient with higher levels of crystallinity, spray-freeze drying allows for the
cooling stage (Duarte, 2016). It can be hypothesized that with further production of porous particles with improved aerodynamic properties
variation of process variables, an increase in purity levels might be and a reduced MMAD.
achieved by a decrease in particle size. Nevertheless, this study
demonstrated the ability to produce pure cocrystal formulations with a 3.5. Comparison of production methods
reduced particle size by the solvent -free spray congealing method.
Carbamazepine-saccharin cocrystal

17
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

The polymorphic carbamazepine-saccharin cocrystal has been pro­ et al. are smaller than that produced by SAS. Supercritical CO2 nano
duced using various techniques, including atomization-based methods spray drying appears to have generated the smallest particles due to
mentioned in this review. Padrela et al. reported polymorphic control of drying aided by both antisolvent and thermal means (Cuadra, 2018;
this cocrystal through the use of several techniques, namely electro­ Padrela et al., 2019).
spraying, supercritical CO2 nano spray drying, and conventional spray Indomethacin-saccharin cocrystal
drying, as well as different solvents (Padrela et al., 2019). Prior to this, The indomethacin-saccharin cocrystal has been produced by several
Cuadra et al. also produced forms I and II of this cocrystal using the methods, including several supercritical atomization-based methods,
supercritical antisolvent (SAS) method (Cuadra, 2018). Cuadra found it namely SAS, AAS (Padrela et al., 2009) and SEA (Padrela et al., 2010).
possible to utilize the SAS method to produce pure cocrystal form I using All three methods were successful in producing the pure cocrystal,
methanol as the solvent at 40 ◦ C, whilst a mixture of forms I and II was regardless of the mechanism of nucleation/crystal growth/solvent
obtained at 60 ◦ C, or while employing ethanol or DMSO as the solvent. removal. The SAS and AAS processes reported no major differences in
The author speculated that changes in temperature and solvent system product composition or particle size with variations in processing con­
affected the ternary phase diagram leading to variations in polymorphic ditions, while no variations were seen in the studying employing the SEA
outcome (Cuadra, 2018). Padrela et al. found solvent selection to have a method (Padrela et al., 2009; Padrela et al., 2010).
negligible effect on polymorphic outcome. Conventional spray drying The resulting particle sizes for SEA produced cocrystal seen in
was incapable of producing a pure polymorphic product, but the use of Fig. 7C, was between 0.3 and 10 μm and were seen as granular and
methanol favoured the production of the metastable form II. Due to the lightly agglomerated. AAS produced cocrystal particles displayed a
relatively large droplets produced during this method, solvent mediated spherical to granular morphology with a particle size between 0.5 and
polymorphic transformation may occur leading to mixtures of poly­ 0.7 μm. The SAS particles however displayed a more elongated shape
morphs. Supercritical CO2 nano spray drying produced form I only, with similar particle sizes to those seen in AAS. Overall, the AAS and SAS
regardless of the solvent employed. This may have been caused by methods produced slightly smaller particle sizes, with the AAS method
several mechanisms within this technique; the antisolvent/spray displaying a more uniform particle size. The CSS method was unsuc­
enhancer role favouring form I, or form I nuclei segregation by nano­ cessful in producing the cocrystal effectively as the reactants are poorly
confinement, or accelerated transformation of form II to I by the rapid soluble within the SCF. For this reason, most experimental conditions
drying process. Opposing to this, electrospraying produced the meta­ produced little to no cocrystal using the CSS method. Cocrystal particles
stable form II of the cocrystal, regardless of solvent choice, possibly due that were produced by this method (seen in Fig. 7D) were large and
to an electrocaloric effect (Padrela et al., 2019). agglomerated, with some particles exceeding 100 μm in length. The
Both studies found that the stable form I displayed a plate like unsuccessfulness of this method can be attributed to the reactants lack of
morphology whilst form II displayed a needle like morphology, but solubility within the SCF, and the exclusion of an atomization step (or
neither study reported any particle sizes. However, from SEM images any other mechanical micronization mechanism) (Padrela et al., 2009).
seen in Fig. 6, it appears as though the cocrystals produced by Padrela

Fig. 6. SEM images of the carbamazepine-saccharin cocrystal produced by electrospraying (form II) (A), supercritical nano spray-drying (SASD/SEA) (form I) (B),
conventional spray drying (form I and II) (C) (reproduced with permissions (Padrela, 2019) and the supercritical antisolvent (SAS) method (form I) (D) (reproduced
with permission (Cuadra, 2018).

18
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

Fig. 7. SEM images of the Indomethacin-Saccharin cocrystal produced by the SAS method (A), AAS method (B) (reproduced with permission (Padrela, 2009), the
SEA method (C) (reproduced with permission (Padrela, 2010) and CSS method (D) (reproduced with permission (Padrela, 2009).

4. Practical considerations cocrystals and salts when administered orally, in that both substances
dissociate, releasing the API. Therefore, cocrystals may follow an
4.1. Regulatory guidelines approval route based on EU Directives 2001/82/EC Article 13(2)(b) and
2001/83/EC Article 10(2)(b) stating that bioequivalence may be
Due to the acceleration of scientific and medical research, it can be demonstrated by appropriate bioavailability studies. Cocrystals which
challenging to effectively introduce regulatory guidelines and proced­ are not administered orally may also fall under this regulatory approval
ures for the continuously expanding list of emerging pharmaceutical route providing there are no differences in safety or efficacy and
technologies, and for new chemical entities as medicines. As a result, appropriate API dissociation is observed. This document also states that
novel technologies may impact and slow down the pace of the regulatory the use of a coformer never before used in a medicinal product, will
approval process, intended to introduce new therapies and medicines for require additional documentation similar to that required for the use of a
patients in a safe manner. In the case of pharmaceutical cocrystals, these new excipient which is considered to be minor, relative to the process for
delays and obstacles were almost magnified to the point that it may not a new chemical entity (NCE) (EMA., 2015).
be a cost-effective opportunity to pursue this type of drug systems. As The FDA published a new draft guidance document on the regulatory
the popularity of publications and patents began to rise (as described in classification of cocrystals in 2016, followed by the publication of the
Fig. 2), the FDA published a draft guidance document on the regulatory final document in 2018. The newest version of this guidance document
classification of cocrystals in 2011. This draft was followed by the (published in 2018) reported that the pharmaceutical cocrystal approval
release of the full guidance document in 2013, describing cocrystals as route will be similar to that of a new API polymorph, and will not be
crystalline solids at room temperature composed of two or more mole­ regarded as a new API or a drug product intermediate (FDA., 2018). For
cules within a crystal lattice. The document recognised the reported pharmaceutical cocrystal approval pathways, two possibilities exist with
similarities between cocrystals, polymorphs and salts, but from a regu­ cocrystals; the new drug application (NDA) pathway (505(b)(2)), and
latory perspective it described pharmaceutical cocrystals as drug prod­ the abbreviated new drug application (ANDA) pathway (505(j)). The
uct intermediates which are to be made subject to additional NDA route is applicable towards cocrystals whose API molecules are not
regulations. This unattractive guidance document led to a plateau in the already a reference listed drug (RLD), which has been the case with
progress of cocrystals in the literature and made them undesirable to the many cocrystals in the market, and thus must contain full reports on the
pharmaceutical industry (Brittain, 2013; FDA, Draft, 2011). In 2015, the safety and efficacy of the drug products. However, if the API molecule
European Medicines Agency (EMA) published a reflective paper within the cocrystal is a previously approved drug (RLD), it may gain
regarding the guidelines towards the use and regulation of cocrystals. approval through the ANDA route, as API polymorphs do. This approval
The EMA described cocrystals as “single phase crystalline structures route focuses on providing evidence that the drug product in question is
made up of two or more components in a definite stoichiometric ratio the same as the RLD regarding administration, dosage, strength and
where the arrangement in the crystal lattice is not based on ionic bonds” conditions of use (FDA., 2018; Research, C.f.D.E.a.2017). For example,
and highlighted that cocrystal components may be neutral or ionized Mayzent, the most recently approved pharmaceutical cocrystal, fol­
(which is discussed in section 2 by molecular and ionic cocrystals). The lowed the NDA approval route as its active ingredient was never
EMA however, outlined a different definition/explanation for pharma­ approved as a standalone drug product (Unger, 2019).
ceutical cocrystals. They stated that an abridged application may be
utilized for cocrystals, which states that they are bioequivalents of an
already approved API. This was justified by the similarity between

19
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

4.2. Quality attributes and particle size distributions adversely affecting the solubility profile of
an API. However, this method adversely introduces various stresses such
In order to determine without doubt that a substance is a cocrystal or as dehydration, shear stresses during atomization, and freezing and
not, for example, a salt, several assumptions listed in the FDAs guidance thawing stresses. Spray flash evaporation is a rapid and continuous
document on cocrystal classification must be proved (FDA., 2018): technique suitable for thermolabile materials, however it requires that
the reactants are both soluble in a low boiling solvent. The field of
• Evidence must be provided to ensure both API and coformer are pharmaceutical cocrystals has been increasing every year through pat­
present within a single unit cell. ents and publications, but the recent updated definition of a cocrystal by
• It must be demonstrated that “substantial dissociation” (FDA., 2018) the FDA in 2018 (see section 4.1) has incentivized pharmaceutical
of API and coformer must occur prior to API interaction and arrival at companies to explore the possibility of generating new cocrystals and
the site of pharmacological activity. introduce a more competitive environment with the ability for quasi-
• Either ΔpKa values or spectroscopic tool must be used to demonstrate generic versions of the API, in the form of cocrystals, to be created.
the production of a cocrystal rather than a salt. If a ΔpKa value < 1 is This new level of competition may be considered as a threat to inno­
achieved, this is sufficient evidence to demonstrate cocrystal vative pharmaceutical companies, as it may prove difficult to argue
production. infringement of a crystal form patent if a cocrystal is produced instead. It
may also serve as an opportunity for these companies to thoroughly
Critical quality attributes of nanocrystals and nano-cocrystals may consider all possible forms prior to a new drug filing, increasing the
differ. Nanocrystals may be designed to follow one of two paths in vivo: degree of scrutiny required by these companies.
solid particle uptake into cells, and fast dissolution allowing for drug
absorption. Due to the nature of cocrystals and their primary function­
CRediT authorship contribution statement
ality generally aiming at improving solubilities, nano-cocrystals follow
the latter path and thus, a set quantity of quality attributes must be
Aaron O’Sullivan: Conceptualization, Writing – original draft.
assessed. These include particle size, particle size distribution, and
Barry Long: Conceptualization, writing - review & editing. Vivek
particle shape and morphology, as well as product compositions and
Verma: Conceptualization, writing - review & editing. Kevin M. Ryan:
solid-state properties. Solid state analysis can be carried out by processes
Supervision, writing - review & editing. Luis Padrela: Conceptualiza­
such as powdered X-ray diffraction (PXRD), differential scanning calo­
tion, Supervision, writing - review & editing.
rimetry (DSC) nuclear magnetic resonance (NMR). Particle size and
shape, which may influence drug solubility, can be examined by dy­
namic light scattering (DLS), and various imaging-based techniques, Declaration of Competing Interest
whilst surface characteristics can be determined by particle zeta po­
tential. Many of these techniques and more will need to be included in The authors declare that they have no known competing financial
the production of pharmaceutical cocrystals to ensure reproducible interests or personal relationships that could have appeared to influence
products with predictable characteristics such as solubility, perme­ the work reported in this paper.
ability, and others (Peltonen, 2018).
Acknowledgements
5. Conclusions
The authors acknowledge Science Foundation Ireland (SFI) for sup­
The production of cocrystals has time and time again proved to porting the work undertaken as part of the SFI frontiers project (grant
effectively improve various physiochemical properties of API molecules 19/FFP/6896), at the SSPC Research Centre (Grant 12/RC/2275_P2),
and the field has continued to expand every year over the last decade. and the financial support provided by Enterprise Ireland (grant CF2017-
The introduction of a nano-sizing step such as atomization further en­ 0754-P).
hances these physiochemical profiles, such as dissolution rates, solubi­
lity, tabletability and permeability. Different atomization-based References
methods have proven effective in producing these nano-cocrystals with
Administration, F.a.D., Caffeine citrate - New drug application (NDA 20-793/S-001. 2000.
each method displaying its own advantages and disadvantages. Super­
Afzal, H., Abbas, N., Hussain, A., Latif, S., Fatima, K., Arshad, M.S., Bukhari, N.I., 2021.
critical fluid-based methods offer a green alternative to more conven­ Physicomechanical, stability, and pharmacokinetic evaluation of aceclofenac
tional methods, such as conventional spray drying, for the production of dimethyl urea cocrystals. AAPS PharmSciTech 22 (2).
nano-sized cocrystals with more narrow particle size distributions than Aitipamula, S., et al., 2012. Polymorphs, salts, and cocrystals: what’s in a name? Cryst.
Growth Des. 12 (5), 2147–2152.
its conventional counterpart. However, these methods are most effective Aitipamula, S., Chow, P.S., Tan, R.B.H., 2010. Polymorphs and Solvates of a Cocrystal
for the removal of miscible organic solvents such as methanol and Involving an Analgesic Drug, Ethenzamide, and 3,5-Dinitrobenzoic Acid. Cryst.
ethanol and encounter issues regarding aqueous solvents. Solvent-based Growth Des. 10 (5), 2229–2238.
Aitipamula, S., Chow, P.S., Tan, R.B.H., 2014. Polymorphism in cocrystals: a review and
SCF methods heavily rely on the solubility of the cocrystal reactants in assessment of its significance. CrystEngComm 16 (17), 3451–3465.
the SCF, which is a constraint in these methods. Although conventional Al-Ani, A.J., et al., 2020. Engineering a New Access Route to Metastable Polymorphs
spray drying typically produces larger micron-sized particles, it has been with Electrical Confinement. Cryst. Growth Des. 20 (3), 1451–1457.
Albertini, B., et al., 2008. New spray congealing atomizer for the microencapsulation of
employed in the pharmaceutical industry for several years and is a much highly concentrated solid and liquid substances. Eur. J. Pharm. Biopharm. 69 (1),
more established method, which implies that the industry possesses a 348–357.
better understanding of the intricacies of this method and may favour it Alhalaweh, A., et al., 2013. Theophylline cocrystals prepared by spray drying:
physicochemical properties and aerosolization performance. AAPS PharmSciTech 14
over novel methodologies. Other methods such as spray freeze drying, (1), 265–276.
spray congealing and spray flash evaporation have been sparingly Alhalaweh, A., Velaga, S.P., 2010. Formation of Cocrystals from Stoichiometric Solutions
employed for cocrystal production but have nevertheless demonstrated of Incongruently Saturating Systems by Spray Drying. Cryst. Growth Des. 10 (8),
3302–3305.
advantages over other methods. Spray congealing offers a green alter­
Al-Otaibi, J.S., Mary, Y.S., Mary, Y.S., Thomas, R., 2019. Quantum mechanical and
native to other methods as it does not require an aqueous or organic photovoltaic studies on the cocrystals of hydrochlorothiazide with isonazid and
phase. This method, however, requires that the API and coformer be malonamide. J. Mol. Struct. 1197, 719–726.
stable at the high temperatures employed. Spray-freeze drying offers Al-Otaibi, J.S., Almuqrin, A.H., Mary, Y.S., Mary, Y.S., Thomas, R., 2020. Cocrystals of
hydrochlorothiazide with picolinamide, tetramethylpyrazine and piperazine:
lower temperatures suitable for thermolabile compounds, particularly quantum mechanical studies, docking and modelling of the photovoltaic efficiency
biological samples, but has been shown to produce larger particle sizes for DSSC. J. Mol. Model. 26 (9).

20
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

Arzi, R.S., Sosnik, A., 2018. Electrohydrodynamic atomization and spray-drying for the Emami, S., 2018. Formulation and physical pharmacy evaluation of cocrystals of
production of pure drug nanocrystals and co-crystals. Adv. Drug Deliv. Rev. 131, biopharmaceutical classification system class 2 drugs. Tabriz University of Medical
79–100. Sciences, Faculty of Pharmacy.
Awasthi, A., M. Dheeraj, H., Birangal, S., Pai, A., Pai, G., Sathyanarayana, M.B., 2021. Emami, S., et al., 2018. Feasibility of electrospray deposition for rapid screening of the
Fabrication of ceritinib cocrystals with improved solubility: Preparation, solid-state cocrystal formation and single step, continuous production of pharmaceutical
characterization, solubility studies, and molecular docking studies. Rasayan J. Chem. nanococrystals. Drug Dev. Ind. Pharm. 44 (6), 1034–1047.
14 (02), 905–913. Emami, F., et al., 2018. Drying Technologies for the Stability and Bioavailability of
Baumann, J.M., Adam, M.S., Wood, J.D., 2021. Engineering Advances in Spray Drying Biopharmaceuticals. Pharmaceutics 10 (3), 131.
for Pharmaceuticals. Annu Rev Chem Biomol Eng 12, 217–240. Fala, L., 2015. Entresto (Sacubitril/Valsartan): First-in-Class Angiotensin Receptor
Bavishi, D.D., Borkhataria, C.H., 2016. Spring and parachute: How cocrystals enhance Neprilysin Inhibitor FDA Approved for Patients with Heart Failure. Am. Health Drug
solubility. Prog. Cryst. Growth Charact. Mater. 62 (3), 1–8. Benefits 8 (6), 330–334.
Bhandari, J., Kanswami, N., Lakshmi, P.K., 2020. Nano Co-crystal Engineering Technique Fatima, K., Bukhari, N.I., Latif, S., Afzal, H., Hussain, A., Shamim, R., Abbas, N., 2021.
to Enhance the Solubility of Ezetimibe. J. Young Pharmacists 12 (2), S10–S15. Amelioration of physicochemical, pharmaceutical, and pharmacokinetic properties
Bitencourt, R.G., et al., 2019. Prediction of solid solute solubility in supercritical CO2 of lornoxicam by cocrystallization with a novel coformer. Drug Dev. Ind. Pharm. 47
with cosolvents using the CPA EoS. Fluid Phase Equilib. 482, 1–10. (3), 498–508.
Bowles, P., Brenek, S.J., Caron, S., Do, N.M., Drexler, M.T., Duan, S., Dubé, P., Hansen, E. FDA, Draft Guidance for Industry on Regulatory Classification of Pharmaceutical Co-
C., Jones, B.P., Jones, K.N., Ljubicic, T.A., Makowski, T.W., Mustakis, J., Nelson, J. Crystals - Food and Drug Administration. Federal Register - National Archives, 2011.
D., Olivier, M., Peng, Z., Perfect, H.H., Place, D.W., Ragan, J.A., Salisbury, J.J., FDA, 2018. Regulatory Classification of Pharmaceutical Co-Crystals Guidance for
Stanchina, C.L., Vanderplas, B.C., Webster, M.E., Weekly, R.M., 2014. Commercial Industry - Food and Drug Administration, U.S.D.o.H.a.H. Services, Editor.
Route Research and Development for SGLT2 Inhibitor Candidate Ertugliflozin. Org. FDA, 2019. Highlights of prescribing information - Reference ID: 4409346 - Food and
Process Res. Dev. 18 (1), 66–81. Drug Administration.
Braga, D., Maini, L., Grepioni, F., 2013. Mechanochemical preparation of co-crystals. Feng, L., Karpinski, P.H., Sutton, P., Liu, Y., Hook, D.F., Hu, B., Blacklock, T.J.,
Chem. Soc. Rev. 42 (18), 7638–7648. Fanwick, P.E., Prashad, M., Godtfredsen, S., Ziltener, C., 2012. LCZ696: a dual-acting
Brittain, H.G., 2013. Pharmaceutical cocrystals: the coming wave of new drug sodium supramolecular complex. Tetrahedron Lett. 53 (3), 275–276.
substances. J. Pharm. Sci. 102 (2), 311–317. Franco, P., De Marco, I., 2020. Supercritical Antisolvent Process for Pharmaceutical
Buddhadev, S.S., Garala, K.C., 2021. Pharmaceutical Cocrystals—A Review. in Applications: A Review. Processes 8 (8), 938.
Multidisciplinary Digital Publishing Institute Proceedings. Franco, P., De Marco, I., 2021. Nanoparticles and Nanocrystals by Supercritical CO2-
Campardelli, R., Oleandro, E., Reverchon, E., 2016. Supercritical assisted injection in a Assisted Techniques for Pharmaceutical Applications: A Review. Applied Sciences 11
liquid antisolvent for PLGA and PLA microparticle production. Powder Technol. 287, (4), 1476.
12–19. Friscic, T., Jones, W., 2009. Recent Advances in Understanding the Mechanism of
Chen, Q., et al., 2018. Experimental and mathematical study of the spray flash Cocrystal Formation via Grinding. Cryst. Growth Des. 9 (3), 1621–1637.
evaporation phenomena. Appl. Therm. Eng. 130, 598–610. Fung, M.H., et al., 2018. Drug-Excipient Interactions: Effect on Molecular Mobility and
Cocrystals, N.-S., 2021. Anti-Tuberculosis Multi-Drug Cocrystals and Nano-Sized Physical Stability of Ketoconazole-Organic Acid Coamorphous Systems. Mol. Pharm.
Cocrystals: Formation Prediction, Polymorph Control and Pharmaceutical 15 (3), 1052–1061.
Advantages. Institute of Pharmaceutical Technology and Biopharmacy Faculty of Fung, M.H., Suryanarayanan, R., 2019. Effect of Organic Acids on Molecular Mobility,
Pharmacy …. Physical Stability, and Dissolution of Ternary Ketoconazole Spray-Dried Dispersions.
Cotabarren, I.M., et al., 2018. Modelling of the spray drying process for particle design. Mol. Pharm. 16 (1), 41–48.
Chem. Eng. Res. Des. 132, 1091–1104. Gallego, R., Bueno, M., Herrero, M., 2019. Sub-and supercritical fluid extraction of
Coty, J.B., et al., 2020. Use of Spray Flash Evaporation (SFE) technology to improve bioactive compounds from plants, food-by-products, seaweeds and microalgae–An
dissolution o poorly soluble drugs: Case study on furosemide nanocrystals. Int. J. update. TrAC, Trends Anal. Chem. 116, 198–213.
Pharm. 589, 119827. Gamidi, R.K., Rasmuson, A.C., 2020. Analysis and Artificial Neural Network Prediction of
Cuadra, I.A., et al., 2016. Pharmaceutical co-crystals of the anti-inflammatory drug Melting Properties and Ideal Mole fraction Solubility of Cocrystals. Cryst. Growth
diflunisal and nicotinamide obtained using supercritical CO2 as an antisolvent. Des. 20 (9), 5745–5759.
J. CO2 Util. 13, 29–37. Ghoreishi, S.M., Hedayati, A., Kordnejad, M., 2016. Micronization of chitosan via rapid
Cuadra, I.A., et al., 2018. Polymorphism in the co-crystallization of the anticonvulsant expansion of supercritical solution. J. Supercrit. Fluids 111, 162–170.
drug carbamazepine and saccharin using supercritical CO2 as an anti-solvent. Ghosh, M., et al., 2020. Preparation of reduced sensitivity co-crystals of cyclic nitramines
J. Supercrit. Fluids 136, 60–69. using spray flash evaporation. Defence Technology 16 (1), 188–200.
Cuadra, I.A., et al., 2020. Cocrystallization of the anticancer drug 5-fluorouracil and Gohel, S.K., Palanisamy, V., Sanphui, P., Prakash, M., Singh, G.P., Chernyshev, V., 2021.
coformers urea, thiourea or pyrazinamide using supercritical CO2 as an antisolvent Isostructural cocrystals of metaxalone with improved dissolution characteristics.
(SAS) and as a solvent (CSS). J. Supercrit. Fluids 160, 104813. RSC Adv. 11 (49), 30689–30700.
Desiraju, G.R., 2013. Crystal engineering: from molecule to crystal. J. Am. Chem. Soc. Gopi, S.P., Banik, M., Desiraju, G.R., 2017. New Cocrystals of Hydrochlorothiazide:
135 (27), 9952–9967. Optimizing Solubility and Membrane Diffusivity. Cryst. Growth Des. 17 (1),
Devarakonda, S.N., 2007. United States Aptent Application Publication - Aripiprazole Co- 308–316.
Crystals. : US 2009/0054455A1. Grossjohann, C., et al., 2015. Polymorphism in sulfadimidine/4-aminosalicylic acid
Dhibar, M., Chakraborty, S., Basak, S., 2021. Assessment of Effects of Solvents on cocrystals: solid-state characterization and physicochemical properties. J. Pharm.
Cocrystallization by Computational Simulation Approach. Curr. Drug Deliv. 18 (1), Sci. 104 (4), 1385–1398.
44–53. Guo, C., Zhang, Q.i., Zhu, B., Zhang, Z., Bao, J., Ding, Q., Ren, G., Mei, X., 2020.
Diniz, L.F., et al., 2021. Multicomponent ionic crystals of diltiazem with dicarboxylic Pharmaceutical Cocrystals of Nicorandil with Enhanced Chemical Stability and
acids toward understanding the structural aspects driving the drug-release. Int. J. Sustained Release. Cryst. Growth Des. 20 (10), 6995–7005.
Pharm. 605, 120790. Ha, E.S., et al., 2020. Pure Trans-Resveratrol Nanoparticles Prepared by a Supercritical
Dizaj, S.M., et al., 2015. Nanosizing of drugs: Effect on dissolution rate. Res. Pharm. Sci. Antisolvent Process Using Alcohol and Dichloromethane Mixtures: Effect of Particle
10 (2), 95–108. Size on Dissolution and Bioavailability in Rats. Antioxidants 9 (4), 342.
do Amaral, L.H.,, et al., 2018. Development and Characterization of Dapsone Cocrystal Halliwell, R.A., et al., 2017. Spray Drying as a Reliable Route to Produce Metastable
Prepared by Scalable Production Methods. AAPS PharmSciTech 19 (6), 2687–2699. Carbamazepine Form IV. J. Pharm. Sci. 106 (7), 1874–1880.
Duarte, I., et al., 2016. Green production of cocrystals using a new solvent-free approach Hanada, S., Fujioka, K., Inoue, Y., Kanaya, F., Manome, Y., Yamamoto, K., 2014. Cell-
by spray congealing. Int. J. Pharm. 506 (1–2), 68–78. Based in Vitro Blood-Brain Barrier Model Can Rapidly Evaluate Nanoparticles’ Brain
Duggirala, N.K., Perry, M.L., Almarsson, Ö., Zaworotko, M.J., 2016. Pharmaceutical Permeability in Association with Particle Size and Surface Modification. Int. J. Mol.
cocrystals: along the path to improved medicines. Chem. Commun. (Camb.) 52 (4), Sci. 15 (2), 1812–1825.
640–655. Harrison, W.T.A., Yathirajan, H.S., Bindya, S., Anilkumar, H.G., Devaraju, 2007.
Eddleston, M.D., et al., 2013. Cocrystallization by Freeze-Drying: Preparation of Novel Escitalopram oxalate: co-existence of oxalate dianions and oxalic acid molecules in
Multicomponent Crystal Forms. Cryst. Growth Des. 13 (10), 4599–4606. the same crystal. Acta Crystallogr. C 63 (2), o129–o131.
Eedara, B.B., Tucker, I.G., Das, S.C., 2018. Cocrystal Approach to Reduce the Aqueous Hasa, D., Schneider Rauber, G., Voinovich, D., Jones, W., 2015. Cocrystal Formation
Solubility and Dissolution Rate for Improved Residence Time of an Anti-Tubercular through Mechanochemistry: from Neat and Liquid-Assisted Grinding to Polymer-
Drug in the Lungs. J. Aerosol Med. Pulmonary Drug Delivery 31 (2), A16. Assisted Grinding. Angewandte Chemie-International Edition 54 (25), 7371–7375.
El-Gizawy, S.A., Osman, M.A., Arafa, M.F., El Maghraby, G.M., 2015. Aerosil as a novel Healy, A.M., Worku, Z.A., Kumar, D., Madi, A.M., 2017. Pharmaceutical solvates,
co-crystal co-former for improving the dissolution rate of hydrochlorothiazide. Int. J. hydrates and amorphous forms: A special emphasis on cocrystals. Adv. Drug Deliv.
Pharm. 478 (2), 773–778. Rev. 117, 25–46.
Elliott, W., Chan, J., 2021. Celecoxib and Tramadol Hydrochloride Tablets (Seglentis) C- Hickey, M.B., Peterson, M.L., Scoppettuolo, L.A., Morrisette, S.L., Vetter, A., Guzmán, H.,
IV. Internal Medicine Alert 43 (22). Remenar, J.F., Zhang, Z., Tawa, M.D., Haley, S., 2007. Performance comparison of a
EMA, Assessment report: Mayzent - European Medicines Agency. Procedure No. EMEA/H/C/ co-crystal of carbamazepine with marketed product. Eur. J. Pharm. Biopharm. 67
004712/0000, 2019. (1), 112–119.
EMA, 2015. Odomzo, sonidegib - European medicines agency. Hiendrawan, S., et al., 2016. Simultaneous cocrystallization and micronization of
EMA, Reflection paper on the use of cocrystals of active substances in medicinal products paracetamol-dipicolinic acid cocrystal by supercritical antisolvent (SAS). Int. J.
- European Medicines Agency. 2015. Pharm. Pharm. Sci. 8, 89–98.
Higashi, K., Ueda, K., Moribe, K., 2017. Recent progress of structural study of
polymorphic pharmaceutical drugs. Adv. Drug Deliv. Rev. 117, 71–85.

21
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

Huang, C., et al., 2018. High-Yielding and Continuous Fabrication of Nanosized CL-20- Lyu, F., et al., 2017. Combined control of morphology and polymorph in spray drying of
Based Energetic Cocrystals via Electrospraying Deposition. Cryst. Growth Des. 18 mannitol for dry powder inhalation. J. Cryst. Growth 467, 155–161.
(4), 2121–2128. MacEachern, L.A., et al., 2021. Ternary Phase Diagram Development and Production of
Imamura, M., et al., 2012. Cocrystal of C-glycoside derivative and L-proline. Google Niclosamide-Urea Co-Crystal by Spray Drying. J. Pharm. Sci. 110 (5), 2063–2073.
Patents. MacEachern, L., Kermanshahi-pour, A., Mirmehrabi, M., 2020. Supercritical Carbon
Jasani, M.S., Kale, D.P., Singh, I.P., Bansal, A.K., 2019. Influence of Drug-Polymer Dioxide for Pharmaceutical Co-Crystal Production. Cryst. Growth Des. 20 (9),
Interactions on Dissolution of Thermodynamically Highly Unstable Cocrystal. Mol. 6226–6244.
Pharm. 16 (1), 151–164. Madan, J.R., Waghmare, S.V., Patil, R.B., Awasthi, R., Dua, K., 2021. Cocrystals of
Jensen, K.T., et al., 2016. Preparation and characterization of spray-dried co-amorphous Apixaban with Improved Solubility and Permeability: Formulation, Physicochemical
drug-amino acid salts. J. Pharm. Pharmacol. 68 (5), 615–624. Characterization, Pharmacokinetic Evaluation, and Computational Studies. Assay
Jindal, A., Singh, R., Tomar, S., Dureja, J., Karan, M., Chadha, R., 2021. Engineering a Drug Dev. Technol. 19 (2), 124–138.
Remedy to Modulate and Optimize Biopharmaceutical Properties of Rebamipide by Maheshwari, C., Jayasankar, A., Khan, N.A., Amidon, G.E., Rodríguez-Hornedo, N.,
Synthesizing New Cocrystal. In Silico and Experimental Studies. Pharm. Res. 38 (12), 2009. Factors that influence the spontaneous formation of pharmaceutical cocrystals
2129–2145. by simply mixing solid reactants. CrystEngComm 11 (3), 493–500.
Jog, R., Burgess, D.J., 2018. Nanoamorphous drug products - Design and development. Mannava, M.K.C., Gunnam, A., Lodagekar, A., Shastri, N.R., Nangia, A.K., Solomon, K.A.,
Int. J. Pharm. 553 (1–2), 238–260. 2021. Enhanced solubility, permeability, and tabletability of nicorandil by salt and
Jung, J., Perrut, M., 2001. Particle design using supercritical fluids: Literature and patent cocrystal formation. CrystEngComm 23 (1), 227–237.
survey. J. Supercrit. Fluids 20 (3), 179–219. Martello, R.H., et al., 2019. Micronization of thymol by RESS and its larvicidal activity
Karashima, M., et al., 2017. Enhanced pulmonary absorption of poorly soluble against Aedes aegypti (Diptera, Culicidae). Ind. Crops Prod. 139, 111495.
itraconazole by micronized cocrystal dry powder formulations. Eur. J. Pharm. Martin, Y.C., 2009. Let’s not forget tautomers. J. Comput. Aided Mol. Des. 23 (10),
Biopharm. 115, 65–72. 693–704.
Karashima, M., Kimoto, K., Yamamoto, K., Kojima, T., Ikeda, Y., 2016. A novel Martin, F., Pop, M., Kacso, I., Grosu, I.G., Miclăuş, M., Vodnar, D., Lung, I., Filip, G.A.,
solubilization technique for poorly soluble drugs through the integration of Olteanu, E.D., Moldovan, R., Nagy, A., Filip, X., Bâldea, I., 2020. Ketoconazole-p-
nanocrystal and cocrystal technologies. Eur. J. Pharm. Biopharm. 107, 142–150. aminobenzoic Acid Cocrystal: Revival of an Old Drug by Crystal Engineering. Mol.
Karimi-Jafari, M., Padrela, L., Walker, G.M., Croker, D.M., 2018. Creating Cocrystals: A Pharm. 17 (3), 919–932.
Review of Pharmaceutical Cocrystal Preparation Routes and Applications. Cryst. Masilamani, S., Ruppelt, S.C., 2003. Escitalopram (Lexapro) for depression. Am. Fam.
Growth Des. 18 (10), 6370–6387. Physician 68 (11), 2235–2236.
Karki, S., Friščić, T., Jones, W., Motherwell, W.D.S., 2007. Screening for pharmaceutical Mathur, V., Satrawala, Y., Rajput, M.S., 2011. Biopharmaceutical perform-ance and
cocrystal hydrates via neat and liquid-assisted grinding. Mol. Pharm. 4 (3), 347–354. stability of co-crystal. Int. J. Pharm. Front. Res. 1 (1), 135–145.
Karmwar, P., et al., 2011. Investigation of properties and recrystallisation behaviour of Matson, D.W., et al., 1987. Rapid Expansion of Supercritical Fluid Solutions - Solute
amorphous indomethacin samples prepared by different methods. Int. J. Pharm. 417 Formation of Powders, Thin-Films, and Fibers. Ind. Eng. Chem. Res. 26 (11),
(1–2), 94–100. 2298–2306.
Kaur, M., Yardley, V., Wang, K.e., Masania, J., Botana, A., Arroo, R.R.J., Li, M., 2021. Matson, D.W., et al., 1987. Rapid expansion of supercritical fluid solutions: solute
Artemisinin Cocrystals for Bioavailability Enhancement. Part 1: Formulation Design formation of powders, thin films, and fibers. Ind. Eng. Chem. Res. 26 (11),
and Role of the Polymeric Excipient. Mol. Pharm. 18 (12), 4256–4271. 2298–2306.
Kavanagh, O.N., Croker, D.M., Walker, G.M., Zaworotko, M.J., 2019. Pharmaceutical Matson, D.W., Petersen, R.C., Smith, R.D., 1986. The preparation of polycarbosilane
cocrystals: from serendipity to design to application. Drug Discov. Today 24 (3), powders and fibers during rapid expansion of supercritical fluid solutions. Mater.
796–804. Lett. 4 (10), 429–432.
Khaw, K.-Y., et al., 2017. Solvent supercritical fluid technologies to extract bioactive Matson, D.W., Petersen, R.C., Smith, R.D., 1987. Production of powders and films by the
compounds from natural sources: A review. Molecules 22 (7), 1186. rapid expansion of supercritical solutions. J. Mater. Sci. 22 (6), 1919–1928.
Kim, W.-J., et al., 2008. Selective caffeine removal from green tea using supercritical Matsuda, Y., et al., 1984. Physicochemical characterization of spray-dried
carbon dioxide extraction. J. Food Eng. 89 (3), 303–309. phenylbutazone polymorphs. J. Pharm. Sci. 73 (2), 173–179.
Kumar, S., 2018. Pharmaceutical cocrystals: an overview. Indian J. Pharm. Sci. 79 (6), Mohammad, M.A., Alhalaweh, A., Velaga, S.P., 2011. Hansen solubility parameter as a
858–871. tool to predict cocrystal formation. Int. J. Pharm. 407 (1–2), 63–71.
Kumar, R., et al., 2021. A critical review on the particle generation and other Mohammady, M., Hadidi, M., Iman Ghetmiri, S., Yousefi, G., 2021. Design of ultra-fine
applications of rapid expansion of supercritical solution. Int. J. Pharm. 608, 121089. carvedilol nanococrystals: Development of a safe and stable injectable formulation.
Kumar, R., et al., 2021. Particle Size Reduction Techniques of Pharmaceutical Eur. J. Pharm. Biopharm. 168, 139–151.
Compounds for the Enhancement of Their Dissolution Rate and Bioavailability. Montes, A., et al., 2016. Precipitation of submicron particles of rutin using supercritical
J. Pharm. Innovation 1–20. antisolvent process. J. Supercrit. Fluids 118, 1–10.
Kuminek, G., Cavanagh, K.L., da Piedade, M.F.M., Rodríguez-Hornedo, N., 2019. Montes, A., et al., 2016. Mangiferin nanoparticles precipitation by supercritical
Posaconazole Cocrystal with Superior Solubility and Dissolution Behavior. Cryst. antisolvent process. J. Supercrit. Fluids 112, 44–50.
Growth Des. 19 (11), 6592–6602. Mujumdar, A.S., Huang, L.X., Chen, X.D., 2010. An overview of the recent advances in
Latif, S., et al., 2021. Improvement of Physico-mechanical and pharmacokinetic spray-drying. Dairy Science & Technology 90 (2–3), 211–224.
attributes of naproxen by cocrystallization with L-alanine. J. Drug Delivery Sci. Mukherjee, A., Desiraju, G.R., 2011. Synthon polymorphism and pseudopolymorphism in
Technol. 61, 102236. co-crystals. The 4,4’-bipyridine-4-hydroxybenzoic acid structural landscape. Chem.
Lefebvre, A.H., McDonell, V.G., 2017. Atomization and sprays. CRC Press. Commun. (Camb.) 47 (14), 4090–4092.
Lennernaäs, H., 1998. Human intestinal permeability. J. Pharm. Sci. 87 (4), 403–410. Mullers, K.C., Paisana, M., Wahl, M.A., 2015. Simultaneous formation and micronization
Lenz, E., et al., 2017. Hot Melt Extrusion and Spray Drying of Co-amorphous of pharmaceutical cocrystals by rapid expansion of supercritical solutions (RESS).
Indomethacin-Arginine With Polymers. J. Pharm. Sci. 106 (1), 302–312. Pharm. Res. 32 (2), 702–713.
Leyssens, T., et al., 2012. Importance of Solvent Selection for Stoichiometrically Diverse Nanoform, What our technology can achieve. 2021.
Cocrystal Systems: Caffeine/Maleic Acid 1:1 and 2:1 Cocrystals. Cryst. Growth Des. Narala, S., et al., 2021. Pharmaceutical Co-Crystals, Salts, and Co-Amorphous Systems: A
12 (3), 1520–1530. Novel Opportunity of Hot Melt Extrusion. J. Drug Deliv. Sci. Technol. 61, 102209.
Li, J.L., et al., 2011. Measurement and correlation of solubility of benzamide in Ndayishimiye, J., Chun, B.S., 2018. Formation, characterization and release behavior of
supercritical carbon dioxide with and without cosolvent. Fluid Phase Equilib. 307 citrus oil-polymer microparticles using particles from gas saturated solutions (PGSS)
(1), 11–15. process. J. Ind. Eng. Chem. 63, 201–207.
Li, H.Q., et al., 2015. Preparation and Performance of Nano HMX/TNT Cocrystals. Negoescu, C.C., et al., 2017. Heat transfer behaviour of supercritical nitrogen in the large
Propellants Explos. Pyrotech. 40 (5), 652–658. specific heat region flowing in a vertical tube. Energy 134, 1096–1106.
Lin, Y., Yang, H., Yang, C., Wang, J., 2014. Preparation, characterization, and evaluation Neurohr, C., et al., 2016. Challenge of the supercritical antisolvent technique SAS to
of dipfluzine-benzoic acid co-crystals with improved physicochemical properties. prepare cocrystal-pure powders of naproxen-nicotinamide. Chem. Eng. J. 303,
Pharm. Res. 31 (3), 566–578. 238–251.
Littringer, E., et al., 2013. Spray drying of aqueous salbutamol sulfate solutions using the Nyström, M., Murtomaa, M., Salonen, J., 2011. Fabrication of amorphous
nano spray dryer B-90—the impact of process parameters on particle size. Drying pharmaceutical materials by electrospraying into reduced pressure. J. Electrostat. 69
Technol. 31 (12), 1346–1353. (4), 351–356.
Liu, H.J., et al., 2021. Bergenin-isonicotinamide (1:1) cocrystal with enhanced solubility O’Nolan, D., Perry, M.L., Zaworotko, M.J., 2016. Chloral hydrate polymorphs and
and investigation of its solubility behavior. J. Drug Delivery Sci. Technol. 64, cocrystal revisited: solving two pharmaceutical cold cases. Cryst. Growth Des. 16 (4),
102556. 2211–2217.
Long, B., et al., 2019. Controlling Polymorphism of Carbamazepine Nanoparticles in a Padrela, L., et al., Cocrystal Screening Using Supercritical Fluid-Assisted Processes. J.
Continuous Supercritical-CO2-Assisted Spray Drying Process. Cryst. Growth Des. 19 Supercritical Fluids The. 53(1): p. 156-164.
(7), 3755–3767. Padrela, L., et al., 2009. Formation of indomethacin-saccharin cocrystals using
Long, B., et al., 2021. Generation and physicochemical characterization of posaconazole supercritical fluid technology. Eur. J. Pharm. Sci. 38 (1), 9–17.
cocrystals using Gas Antisolvent (GAS) and Supercritical Solvent (CSS) methods. Padrela, L., et al., 2010. Screening for pharmaceutical cocrystals using the supercritical
J. Supercrit. Fluids 170, 105134. fluid enhanced atomization process. J. Supercrit. Fluids 53 (1–3), 156–164.
Lu, W.X., et al., 2022. Unveiling the importance of process parameters on droplet Padrela, L., et al., 2014. Tuning physicochemical properties of theophylline by
shrinkage and crystallization behaviors of easily crystalline material during spray cocrystallization using the supercritical fluid enhanced atomization technique.
drying. Drying Technol. 40 (2), 326–336. J. Supercrit. Fluids 86, 129–136.

22
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

Padrela, L., et al., 2018. Supercritical carbon dioxide-based technologies for the Ren, B.-Y., Dai, X.-L., Wang, J., Wu, C., Lu, T.-B., Chen, J.-M., 2021. Cocrystallization of
production of drug nanoparticles/nanocrystals – A comprehensive review. Adv. Drug axitinib with carboxylic acids: preparation, crystal structures and dissolution
Deliv. Rev. 131, 22–78. behavior. CrystEngComm 23 (32), 5504–5515.
Padrela, L., et al., 2018. Supercritical carbon dioxide-based technologies for the Research, C.f.D.E.a., Determining Whether to Submit an ANDA or a 505(b)(2) Application.
production of drug nanoparticles/nanocrystals - A comprehensive review. Adv. Drug FDA - Guidance document, 2017.
Deliv. Rev. 131, 22–78. Reverchon, E., De Marco, I., 2011. Mechanisms controlling supercritical antisolvent
Padrela, L.M., et al., 2019. Co-crystal polymorphic control by nanodroplet and electrical precipitate morphology. Chem. Eng. J. 169 (1–3), 358–370.
confinement. CrystEngComm 21 (18), 2845–2848. Ribas, M.M., et al., 2019. Curcumin-nicotinamide cocrystallization with supercritical
Paisana, M.C., et al., 2016. Production and stabilization of olanzapine nanoparticles by solvent (CSS): Synthesis, characterization and in vivo antinociceptive and anti-
rapid expansion of supercritical solutions (RESS). J. Supercrit. Fluids 109, 124–133. inflammatory activities. Ind. Crops Prod. 139, 111537.
Pan, Y.J., et al., 2020. Supercritical antisolvent process-assisted fabrication of chrysin- Rodrigues, M.A., et al., 2009. Anti-solvent effect in the production of lysozyme
polyvinylpyrrolidone sub-microparticles for improved anticancer efficiency. nanoparticles by supercritical fluid-assisted atomization processes. J. Supercrit.
J. Supercrit. Fluids 162, 104847. Fluids 48 (3), 253–260.
Pantwalawalkar, J., et al., 2021. Novel curcumin ascorbic acid cocrystal for improved Rodrigues, M.A., et al., 2012. Development of a novel mucosal vaccine against strangles
solubility. J. Drug Delivery Sci. Technol. 61, 102233. by supercritical enhanced atomization spray-drying of Streptococcus equi extracts
Panzade, P., Shendarkar, G., Kulkarni, D., Shelke, S., 2021. Solid State Characterization and evaluation in a mouse model. Eur. J. Pharm. Biopharm. 82 (2), 392–400.
and Dissolution Enhancement of Nevirapine Cocrystals. Adv. Pharm. Bull. 11 (4), Rodrigues, M., et al., 2018. Pharmaceutical cocrystallization techniques. Advances and
772–776. challenges. Int J Pharm 547 (1–2), 404–420.
Park, H.J., et al., 2007. Recrystallization of fluconazole using the supercritical Rodriguez-Aller, M., Guillarme, D., Veuthey, J.-L., Gurny, R., 2015. Strategies for
antisolvent (SAS) process. Int. J. Pharm. 328 (2), 152–160. formulating and delivering poorly water-soluble drugs. J. Drug Delivery Sci.
Patel, R., Patel, M., Suthar, A., 2009. Spray drying technology: an overview. Indian J. Sci. Technol. 30, 342–351.
Technol. 2 (10), 44–47. Roy, P., Ghosh, A., 2020. Progress on cocrystallization of poorly soluble NME’s in the last
Patel, R.D., Raval, M.K., 2020. Formulation of Diacerein Cocrystal Using β-Resorcylic decade. CrystEngComm 22 (42), 6958–6974.
Acid for Improvement of Physicomechanical and Biopharmaceutical Properties. Org. Ruhan, A., Motonobu, G., 2009. Supercritical fluid extraction in food analysis. Handbook of
Process Res. Dev. 25 (3), 384–394. food analysis instruments. CRC Press, Taylor and Francis, Boca Raton, FL, pp. 25–54.
Patil, S., et al., 2018. Electrosprayed forskolin cocrystals with enhanced aqueous Said-Galiyev, E., Pototskaya, I., Vygodskii, Y.S., 2004. Supercritical carbon dioxide and
solubility. Analytical Chemistry Letters 8 (3), 321–330. polymers. Polym. Sci., Series C: Rev. 46 (1), 1–13.
Patil, S., Kulkarni, J., Mahadik, K., 2016. Exploring the Potential of Electrospray Salem, A., Nagy, S., Pál, S., Széchenyi, A., 2019. Reliability of the Hansen solubility
Technology in Cocrystal Synthesis. Ind. Eng. Chem. Res. 55 (30), 8409–8414. parameters as co-crystal formation prediction tool. Int. J. Pharm. 558, 319–327.
Patil, S., Xalxo, K., Mahadik, K., 2017. Probing Influence of Solvent on Polymorphic Salem, A., Takácsi-Nagy, A., Nagy, S., Hagymási, A., Gősi, F., Vörös-Horváth, B., Balić, T.,
Transformation of Carbamazepine Using Electrospray Technology. Pál, S., Széchenyi, A., 2021. Synthesis and Characterization of Nano-Sized 4-Ami­
J. Pharmaceutical Innovation 12 (4), 309–318. nosalicylic Acid-Sulfamethazine Cocrystals. Pharmaceutics 13 (2), 277.
Patil, S., Ujalambkar, V., Mahadik, A., 2017. Electrospray technology as a probe for Sander, J.R.G., Bučar, D.-K., Henry, R.F., Zhang, G.G.Z., MacGillivray, L.R., 2010.
cocrystal synthesis: Influence of solvent and coformer structure. J. Drug Delivery Sci. Pharmaceutical Nano-Cocrystals: Sonochemical Synthesis by Solvent Selection and
Technol. 39, 217–222. Use of a Surfactant. Angewandte Chemie-International Edition 49 (40), 7284–7288.
Patil, S., Chaudhari, K., Kamble, R., 2018. Electrospray technique for cocrystallization of Sanphui, P., Devi, V.K., Clara, D., Malviya, N., Ganguly, S., Desiraju, G.R., 2015.
phytomolecules. J. King Saud Univ. Sci. 30 (1), 138–141. Cocrystals of Hydrochlorothiazide: Solubility and Diffusion/Permeability
Pekamwar, S.S., Kulkarni, D.A., 2021. DEVELOPMENT AND EVALUATION OF Enhancements through Drug-Coformer Interactions. Mol. Pharm. 12 (5), 1615–1622.
BICOMPONENT COCRYSTALS OF ACECLOFENAC FOR EFFICIENT DRUG Sathisaran, I., Dalvi, S.V., 2018. Engineering cocrystals of poorly water-soluble drugs to
DELIVERY WITH ENHANCED SOLUBILITY AND IMPROVED DISSOLUTION. Indian enhance dissolution in aqueous medium. Pharmaceutics 10 (3), 108.
Drugs 58 (08), 54–60. Schittny, A., Huwyler, J., Puchkov, M., 2020. Mechanisms of increased bioavailability
Peltonen, L., 2018. Practical guidelines for the characterization and quality control of through amorphous solid dispersions: a review. Drug Delivery 27 (1), 110–127.
pure drug nanoparticles and nano-cocrystals in the pharmaceutical industry. Adv. Schmelzer, J.W., Abyzov, A.S., 2017. How do crystals nucleate and grow: Ostwald’s rule
Drug Deliv. Rev. 131, 101–115. of stages and beyond. In: Thermal physics and thermal analysis. Springer,
Peng, H.H., et al., 2019. Preparation of pH-responsive DOX-loaded chitosan pp. 195–211.
nanoparticles using supercritical assisted atomization with an enhanced mixer. Int. Serrano, D.R., et al., 2016. Modelling and shadowgraph imaging of cocrystal dissolution
J. Pharm. 558, 82–90. and assessment of in vitro antimicrobial activity for sulfadimidine/4-aminosalicylic
Pessi, J., et al., 2016. Controlled Expansion of Supercritical Solution: A Robust Method to acid cocrystals. Eur. J. Pharm. Sci. 89, 125–136.
Produce Pure Drug Nanoparticles With Narrow Size-Distribution. J. Pharm. Sci. 105 Serrano, D.R., et al., 2016. Cocrystal habit engineering to improve drug dissolution and
(8), 2293–2297. alter derived powder properties. J. Pharm. Pharmacol. 68 (5), 665–677.
Pessoa, A.S., et al., 2019. Precipitation of resveratrol-isoniazid and resveratrol- Setyawan, D., Adyaksa, F.R., Sari, H.L., Paramita, D.P., Sari, R., 2021. Cocrystal
nicotinamide cocrystals by gas antisolvent. J. Supercrit. Fluids 145, 93–102. formation of loratadine-succinic acid and its improved solubility. J. Basic Clin.
Petersen, R.C., Matson, D.W., Smith, R.D., 1986. Rapid precipitation of low vapor Physiol. Pharmacol. 32 (4), 623–630.
pressure solids from supercritical fluid solutions: the formation of thin films and Shaikh, R., Singh, R., Walker, G.M., Croker, D.M., 2018. Pharmaceutical Cocrystal Drug
powders. J. Am. Chem. Soc. 108 (8), 2100–2102. Products: An Outlook on Product Development. Trends Pharmacol. Sci. 39 (12),
Petrow, V., A.J., Stephenson, T., Stephenson, O., 1962. COMPOUND OF BETAINE AND 1033–1048.
CHILORAL AND METHOD FOR PREPARING SAME. US Patent Office. Shattock, T.R., et al., 2008. Hierarchy of Supramolecular Synthons: Persistent Carboxylic
Petruševski, G., Naumov, P., Jovanovski, G., Ng, S.W., 2008. Unprecedented Acid center dot center dot center dot Pyridine Hydrogen Bonds in Cocrystals That
sodium–oxygen clusters in the solid-state structure of trisodium also Contain a Hydroxyl Moiety. Cryst. Growth Des. 8 (12), 4533–4545.
hydrogentetravalproate monohydrate: A model for the physiological activity of the Shinozaki, T., Ono, M., Higashi, K., Moribe, K., 2019. A Novel Drug-Drug Cocrystal of
anticonvulsant drug Epilim®. Inorg. Chem. Commun. 11 (1), 81–84. Levofloxacin and Metacetamol: Reduced Hygroscopicity and Improved
Pi, J., Wang, S., Li, W., Kebebe, D., Zhang, Y., Zhang, B., Qi, D., Guo, P., Li, N., Liu, Z., Photostability of Levofloxacin. J. Pharm. Sci. 108 (7), 2383–2390.
2019. A nano-cocrystal strategy to improve the dissolution rate and oral Smith, R.D., et al., 1986. Performance of capillary restrictors in supercritical fluid
bioavailability of baicalein. Asian J. Pharm. Sci. 14 (2), 154–164. chromatography. Anal. Chem. 58 (9), 2057–2064.
Piatkowski, M., Zbicinski, I., 2006. Analysis of the mechanism of counter-current spray Smith, A.J., Kim, S.-H., Duggirala, N.K., Jin, J., Wojtas, L., Ehrhart, J., Giunta, B., Tan, J.,
drying. In: Drying of Porous Materials. Springer, pp. 89–101. Zaworotko, M.J., Shytle, R.D., 2013. Improving lithium therapeutics by crystal
Preiss, F.J., Hetz, M., Karbstein, H.P., 2022. Does Cavitation Affect Droplet Breakup in engineering of novel ionic cocrystals. Mol. Pharm. 10 (12), 4728–4738.
High-Pressure Homogenization? Chem. Ing. Tech. 94 (3), 374–384. Sodeifian, G., Sajadian, S.A., Daneshyan, S., 2018. Preparation of Aprepitant
Qiao, N., Li, M., Schlindwein, W., Malek, N., Davies, A., Trappitt, G., 2011. nanoparticles (efficient drug for coping with the effects of cancer treatment) by rapid
Pharmaceutical cocrystals: an overview. Int. J. Pharm. 419 (1-2), 1–11. expansion of supercritical solution with solid cosolvent (RESS-SC). J. Supercrit.
Ranjan, S., Devarapalli, R., Kundu, S., Vangala, V.R., Ghosh, A., Reddy, C.M., 2017. Fluids 140, 72–84.
Three new hydrochlorothiazide cocrystals: Structural analyses and solubility studies. Sökmen, M., Demir, E., Alomar, S.Y., 2018. Optimization of sequential supercritical fluid
J. Mol. Struct. 1133, 405–410. extraction (SFE) of caffeine and catechins from green tea. J. Supercritical Fluids 133,
Rao, S., et al., 2011. Particle size reduction to the nanometer range: a promising 171–176.
approach to improve buccal absorption of poorly water-soluble drugs. Int. J. Solaimalai, R., et al., 2019. Synthesis of 5-hydroxy-2-methyl-naphthalene-1,4-dione
Nanomed. 6, 1245–1251. cocrystals with pyridine-3-carboxamide using electrospray technology:
Ray, E., et al., 2020. Autophagy-Inducing Inhalable Co-crystal Formulation of physicochemical characterization and in vitro non-everted rat intestinal absorption
Niclosamide-Nicotinamide for Lung Cancer Therapy. AAPS PharmSciTech 21 (7), study. New J. Chem. 43 (15), 5687–5696.
260. Sovago, I., et al., 2016. Properties of the Sodium Naproxen-Lactose-Tetrahydrate Co-
Reddy, D.S., Ovchinnikov, Y.E., Shishkin, O.V., Struchkov, Y.T., Desiraju, G.R., 1996. Crystal upon Processing and Storage. Molecules 21 (4), 509.
Supramolecular Synthons in Crystal Engineering. 3. Solid State Architecture and Spitzer, D., et al., 2014. Continuous engineering of nano-cocrystals for medical and
Synthon Robustness in Some 2,3-Dicyano-5,6-dichloro-1,4-dialkoxybenzenes 1. energetic applications. Sci. Rep. 4 (1), 6575.
J. Am. Chem. Soc. 118 (17), 4085–4089. Sun, J., et al., 2012. Effect of particle size on solubility, dissolution rate, and oral
Reid, R.C., Prausnitz, J.M., Poling, B.E., 1987. The properties of gases and liquids. bioavailability: evaluation using coenzyme Q10 as naked nanocrystals. Int. J.
Nanomed. 7, 5733.

23
A. O’Sullivan et al. International Journal of Pharmaceutics 621 (2022) 121798

Tan, J.B., Liu, J.H., Ran, L.L., 2021. A Review of Pharmaceutical Nano-Cocrystals: A Wang, W.X., et al., 2021. A cocrystal for effectively reducing the hepatotoxicity of
Novel Strategy to Improve the Chemical and Physical Properties for Poorly Soluble ethionamide. J. Mol. Struct. 1243, 130729.
Drugs. Crystals 11 (5), 463. Wang, B.C., Su, C.S., 2020. Solid solubility measurement of ipriflavone in supercritical
Tanaka, R., et al., 2020. Application of spray freeze drying to theophylline-oxalic acid carbon dioxide and microparticle production through the rapid expansion of
cocrystal engineering for inhaled dry powder technology. Drug Dev. Ind. Pharm. 46 supercritical solutions process. J. CO2 Util. 37, 285–294.
(2), 179–187. Wang, F.-Y., Zhang, Q.i., Zhang, Z., Gong, X., Wang, J.-R., Mei, X., 2018. Solid-state
Tawfeek, H.M., Chavan, T., Kunda, N.K., 2020. Effect of Spray Drying on Amorphization characterization and solubility enhancement of apremilast drug-drug cocrystals.
of Indomethacin Nicotinamide Cocrystals; Optimization, Characterization, and CrystEngComm 20 (39), 5945–5948.
Stability Study. AAPS PharmSciTech 21 (5), 181. Weng, J.W., et al., 2019. Cocrystal Engineering of Itraconazole with Suberic Acid via
T-Cells, C., Novartis receives positive CHMP opinion for Mayzent®(siponimod) for the Rotary Evaporation and Spray Drying. Cryst. Growth Des. 19 (5), 2736–2745.
treatment of adult patients with active secondary progressive multiple sclerosis (SPMS). Weyna, D.R., Shattock, T., Vishweshwar, P., Zaworotko, M.J., 2009. Synthesis and
Thimmasetty, J., Ghosh, T., Nayak, N.S., Raheem, A., 2021. Oral Bioavailability Structural Characterization of Cocrystals and Pharmaceutical Cocrystals:
Enhancement of Paliperidone by the use of Cocrystalization and Precipitation Mechanochemistry vs Slow Evaporation from Solution. Cryst. Growth Des. 9 (2),
Inhibition. J. Pharm. Innovation 16 (1), 160–169. 1106–1123.
Tiago, J.M., et al., 2013. Single-Step Co-Crystallization and Lipid Dispersion by Wicaksono, Y., et al., 2021. Preparation of Spray Dried Coamorphous Solids to Improve
Supercritical Enhanced Atomization. Cryst. Growth Des. 13 (11), 4940–4947. the Solubility and Dissolution Rate of Atorvastatin Calcium. Jurnal Teknologi-
Tjandrawinata, R.R., Hiendrawan, S., Veriansyah, B., 2019. Processing Paracetamol-5- Sciences & Engineering 83 (2), 77–83.
Nitroisophthalic Acid Cocrystal Using Supercritical CO2 as an Anti-Solvent. Int. J. Witika, B.A., Smith, V.J., Walker, R.B., 2021. Top-Down Synthesis of a Lamivudine-
Appl. Pharm. 11, 194–199. Zidovudine Nano Co-Crystal. Crystals 11 (1), 33.
Türk, M., Bolten, D., 2016. Polymorphic properties of micronized mefenamic acid, Wu, W.Y., Su, C.S., 2018. Recrystallization and Production of Spherical Submicron
nabumetone, paracetamol and tolbutamide produced by rapid expansion of Particles of Sulfasalazine Using a Supercritical Antisolvent Process. Crystals 8 (7),
supercritical solutions (RESS). The J. Supercritical Fluids 116, 239–250. 295.
Unger, E., 2019. NDA approval letter 209884, from the Department of Health and Xiang, S.T., et al., 2019. Solubility measurement and RESOLV-assisted nanonization of
Human Services. gambogic acid in supercritical carbon dioxide for cancer therapy. J. Supercrit. Fluids
Upadhye, K.P., et al., Development and Evaluation of Taste Masked Azithromycin by 150, 147–155.
Crystal Engineering. Xu, P.Y., et al., 2020. Supercritical carbon dioxide-assisted nanonization of
Urano, M., et al., 2020. Physical Characteristics of Cilostazol-Hydroxybenzoic Acid dihydromyricetin for anticancer and bacterial biofilm inhibition efficacies.
Cocrystals Prepared Using a Spray Drying Method. Crystals 10 (4), 313. J. Supercrit. Fluids 161, 104840.
Vaksler, Y.A., et al., 2021. Spectroscopic characterization of single co-crystal of Yoon, T.J., et al., 2016. Tetracycline nanoparticles precipitation using supercritical and
mefenamic acid and nicotinamide using supercritical CO2. J. Mol. Liq. 334, 116117. liquid CO2 as antisolvents. J. Supercrit. Fluids 107, 51–60.
Van Eerdenbrugh, B., Vermant, J., Martens, J.A., Froyen, L., Humbeeck, J.V., Van den Yu, G., et al., 2021. Study on solubilization of telmisartan by forming cocrystal with
Mooter, G., Augustijns, P., 2010. Solubility increases associated with crystalline drug aromatic carboxylic acid. CrystEngComm.
nanoparticles: methodologies and significance. Mol. Pharm. 7 (5), 1858–1870. Yu, Y.-M., Niu, Y.-Y., Wang, L.-Y., Li, Y.-T., Wu, Z.-Y., Yan, C.-W., 2021. Supramolecular
Vemavarapu, C., Mollan, M.J., Needham, T.E., 2009. Coprecipitation of pharmaceutical self-assembly and perfected in vitro/vivo property of 5-fluorouracil and ferulic acid
actives and their structurally related additives by the RESS process. Powder Technol. on the strength of double optimized strategy: the first 5-fluorouracial-phenolic acid
189 (3), 444–453. nutraceutical cocrystal with synergistic antitumor efficacy. Analyst 146 (8),
Vemuri, V.D., Lankalapalli, S., 2021. Amino acid based Rosuvastatin cocrystals: Towards 2506–2519.
the improvement of physicochemical parameters. J. Cryst. Growth 570, 126241. Yu, L., 2001. Amorphous pharmaceutical solids: preparation, characterization and
Vorobei, A.M., Parenago, O.O., 2021. Using Supercritical Fluid Technologies to Prepare stabilization. Adv. Drug Deliv. Rev., 2001. 48(1): p. 27-42.
Micro- and Nanoparticles. Russ. J. Phys. Chem. A 95 (3), 407–417. Zhang, Y.-X., Wang, L.-Y., Dai, J.-K., Liu, F., Li, Y.-T., Wu, Z.-Y., Yan, C.-W., 2019. The
Walsh, D., et al., 2018. Production of cocrystals in an excipient matrix by spray drying. comparative study of cocrystal/salt in simultaneously improving solubility and
Int. J. Pharm. 536 (1), 467–477. permeability of acetazolamide. J. Mol. Struct. 1184, 225–232.
Walsh, D., et al., 2018. Engineering of pharmaceutical cocrystals in an excipient matrix: Zhao, Z.Y., et al., 2018. Co-Crystal of Paracetamol and Trimethylglycine Prepared by a
Spray drying versus hot melt extrusion. Int. J. Pharm. 551 (1–2), 241–256. Supercritical CO2 Anti-Solvent Process. Chem. Eng. Technol. 41 (6), 1122–1131.

24

You might also like