Next Article in Journal
The Microstructures and Mechanical Properties of a Welded Ni-Based Hastelloy X Superalloy
Next Article in Special Issue
(Z)-2-(2-(Anthracen-9-ylmethylene)hydrazineyl)pyrimidine-rhodium(III): Crystal Structure and DNA Binding Property
Previous Article in Journal
Synthesis and Solid-State X-ray Structure of the Mononuclear Palladium(II) Complex Based on 1,2,3-Triazole Ligand
Previous Article in Special Issue
Experimental and Hirshfeld Surface Investigations for Unexpected Aminophenazone Cocrystal Formation under Thiourea Reaction Conditions via Possible Enamine Assisted Rearrangement
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cocrystals and Drug–Drug Cocrystals of Anticancer Drugs: A Perception towards Screening Techniques, Preparation, and Enhancement of Drug Properties

by
Divya Dhatri Kara
and
Mahalaxmi Rathnanand
*
Department of Pharmaceutics, Manipal College of Pharmaceutical Sciences, Manipal Academy of Higher Education, Manipal 576104, Karnataka, India
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(10), 1337; https://doi.org/10.3390/cryst12101337
Submission received: 16 August 2022 / Revised: 10 September 2022 / Accepted: 13 September 2022 / Published: 21 September 2022

Abstract

:
The most favored approach for drug administration is the oral route. Several anticancer drugs come under this category and mostly lack solubility and oral bioavailability, which are the most common causes of inadequate clinical efficiency. Enhancing oral absorption of anticancer drugs with low aqueous solubility and drug impermeability is currently an effective area of research. Many scientists have looked into pharmaceutical cocrystals as a way to improve the physicochemical properties of several anticancer drugs. Benefits of pharmaceutical cocrystals over other solid forms may include improved solubility, bioavailability, and a reduced susceptibility for phase transition. Cocrystal strategy also stands as a green synthesis tool by using very limited organic solvents during its formulation. Having so many advantages, to date, the reported cocrystals and drug–drug cocrystals of anticancer drugs are limited. Here we review the pharmaceutical cocrystals and drug–drug cocrystals of the anticancer drugs reported in the last decade and their future in imaging, and also shed light on the opportunities and challenges for the development of anticancer drug cocrystals.

1. Introduction

Cancer is regarded as one of the prevailing and major reasons for increased death worldwide and creating a huge global public health concern [1]. Cancer is one of the leading reasons for death, standing second globally with one in six deaths [2]. In accordance with the Global Cancer Observatory (GCO), an estimated nineteen million new instances and ten million deaths from cancer occurred in 2020 [3]. Even though these impressive numbers are testimony that the “war on cancer” has not been won [4], by the next decade, the projected global load is likely to rise to three billion additional cases and sixteen million deaths from cancer [5].
There are different types of cancers out of which a few occurring kinds are colorectal cancer, endometrial cancer, lung cancer, kidney cancer, bladder cancer, leukaemia, pancreatic cancer, melanoma, prostate cancer, breast cancer, non-Hodgkin lymphoma, uterine cancer, skin cancer, liver cancer, thyroid cancer [6]. The facts and statistics of several cancers are depicted in Figure 1.
Over the last decade, due to intensified prevalence of cancer, it has been a challenge for health workers [7]. The discipline of science and technology has continuously been working on gaps in its field and witnessed tremendous relative growth in filling the gaps related to drug therapies [8]. Until the previous few decades, most of the cancer treatment was administered solely through the parenteral route. Nevertheless, considering the ease and frequency of dose administration, when compared to the parenteral route, the oral route has attracted a lot of attention [9].
The oral route of administration is one of the most common and conventional methods for drug delivery; with the major benefits of improved safety, comfort, and patient acceptance [10]. The ability to create customized formulations based on the physicochemical attributes of drug substances stands appealing to the scientific community. However, due to extensive first-pass effect, lower aqueous solubility, higher efflux ratio, impermeability, and bioavailability, oral administration of cancer chemotherapeutic agents is hampered [11].
About 75% of new chemical entities (NCEs), notably anticancer drugs, are believed to be highly water-insoluble molecules [12]. Aqueous solubility stands as a major roadblock to the development and clinical application of these anticancer agents. “According to the Biopharmaceutics Classification System (BCS), a drug is considered to be poorly water-soluble if its highest dose strength is not soluble in 250 mL or less of aqueous media over the pH range of 1 to 7.4” [13]. Unfortunately, many anticancer drugs fall into BCS class 2 and 4 with low solubility, for example, curcumin [14], paclitaxel [15], dasatinib [16], docetaxel [17] and many have a promising antineoplastic activity, which falls into this category. These drugs usually have high lattice energies which restrict the drug from dissolving and form hydrogen bonds with water [18], further restricting the solubility typically below the microgram per milliliter range.
Over the past few decades, many efforts have been made to resolve the lower solubility issues and also improve the drug properties via various novel formulations such as amorphous solid dispersions (ASDs) [19], complexation [20], lipid-based systems [21], nanoparticles [22], and crystal engineering [23], as well as the application of new technologies such as hot-melt extrusion [24], spray-drying [25] and other hybridization technologies [26] to circumvent solubility, dissolution rate, and bioavailability problems of poorly aqueous soluble drugs [27]. Each of these formulation systems has its cons of instability, moisture uptake, poor scale-up, high methodology cost, thermal instability, and other excipient incompatibilities. Especially, though ASDs success rate in enhancing the solubility is higher than the cocrystals, their lack of stability and recrystallization always stands as a problem that clears way to form cocrystals when there is no such point for recrystallization.
Since oral administration is the most convenient route of administration; solid dosage forms, such as tablets and capsules, account for more than 80% of commercial pharmaceuticals [28]. Polymorphs, amorphous forms, cocrystals, salts, and their hydrates can be employed to change the physicochemical properties of drugs for solid-state formulations [29]. The processing, distribution, and therapeutic effectiveness of any dosage form are all influenced by the physicochemical parameters of active pharmaceutical ingredients (APIs), such as stability, dissolution, hygroscopicity, and solubility [30]. There is still so much scope for the breakthrough of desirable stable forms of an API for clinical application by changing just the physicochemical properties without disrupting the molecular-level chemical form of the API where cocrystals and salts are the best suited for justification.
The most promising method for increasing API solubility and dissolution is salt formation [31]. Salts, on the other hand, have the disadvantage of being more likely to form hydrates over cocrystals, and many have a high hygroscopicity, rendering them susceptible to deliquescence upon water sorption [32].
Cocrystallization has increased interest in recent years, and as mentioned in Figure 2, proves to be one of the approaches to improve the bioavailability of water-insoluble drugs [33].
Unlike salts, non-ionizable drugs can benefit from cocrystallization [34]. Cocrystals do not have the physicochemical stability concerns that ASDs do due to their crystalline form as mentioned in Figure 3 [35].
In cocrystallization, with several advantages come various disadvantages, such as difficulty in scaling up, efficiency problems, may form solvates, issues during stability such as hydrolysis and not being appropriate for thermolabile drugs.
The main objective of this review is to highlight recent breakthroughs in enhancing the solubility, oral bioavailability, and efficacy of anticancer drugs via cocrystallization. Cocrystals of different anticancer drugs are studied in terms of their design, discovery, preparation, and characterization to demonstrate how these new crystalline forms can improve APIs characteristics. Additionally, new aspects of green synthesis and the advantages of imaging platforms through cocrystals have been discussed. Finally, we review the research on cocrystals of numerous anticancer drugs, concentrating on how many cocrystal frameworks could improve pharmacokinetic behavior.

2. Cocrystals for the Pharmaceutical Formulation Development

Pharmaceutical cocrystals are crystalline compounds that have two or more different molecules in their crystal lattice [36]. The benefits of cocrystals include their stable crystalline shape and do not include any other excipients [37]. The attributes that play a significant part in influencing physicochemical characteristics include the properties of APIs and coformers [38].
As mentioned, pharmaceutical cocrystals may have several distinct advantages over all the other orally administered solid dosage forms and allow the highly efficacious small molecules that may disappear if all their drawbacks are not rectified [39]. Adding more light to its pharmaceutical cocrystals provides more prospects to the investigators by improving the lifespan of the dosage form as a part of its intellectual properties [40].
The low solubility of drugs in an aqueous medium leads to diminished bioavailability [41]. In this regard, several formulation techniques are available for enhancing the drug solubility or dissolution by their enhancement mechanisms, such as particle technology for increasing drug solubility using micronization, nanocarriers, crystal engineering, etc., [42]. As mentioned, cocrystals follow the spring and parachute effect, which enables the use of supersaturation as a strategy to increase the solubility and dissolution rate of poorly soluble APIs; however, two crucial processes must be maintained:
  • Formation of the supersaturated metastable state.
  • Maintaining the above condition.
Supersaturation of poorly water-soluble APIs can be achieved using a variety of approaches, which includes crystalline salts, cocrystals, ASD in polymers, metallic frameworks, and nanoparticles.
In general terms, drug bioavailability is directly proportional to the dissolution [43]. Most of the techniques such as ASDs, eutectics, and cocrystals follow the conceptual mechanism of the “spring and parachute effect” [44].
(1)
The solubility of the crystalline (stable) state is low.
(2)
The metastable form (amorphous) has a maximum solubility but rapidly decreases to the poorly soluble form (crystalline form).
Details regarding the “spring and parachute” effect are depicted in Figure 4, which explains the phenomenon of the “spring effect” where the drug dissolves immediately, creating a super saturation concentration, and most significantly, the supersaturated state, which is thermodynamically unstable has to persist for a comparatively extended time to ensure that the drug is absorbed effectively; this phenomenon is identified as the “parachute effect” [45]. In case the “parachute period” is rapid, the drug in its “spring stage” will precipitate quickly, resulting in reduced drug absorption [46]. As a result, the benefits of increased dissolution during the spring effect would be reduced, and simultaneously the APIs availability [47]. To constrict this precipitation and extend the parachute effect and avoid the precipitation of the unstable crystal forms, hydrophilic polymers such as HPMC, PVP, etc., can be added to inhibit precipitation [48]. They also add in synergism, which improves the dissolution, extends the parachute effect, and proportionally improves the absorption and bioavailability of cocrystals [49].

3. Coformer Screening and Selection Techniques

Selecting the best suitable coformer stands as a crucial phase to accessing the cocrystals with suitable properties and simultaneously crystal engineering of APIs is a difficult task [50].
Since the range of coformers that should be evaluated often reaches hundreds and more than one method is frequently used to validate cocrystal generation, this makes the selection of acceptable coformers the most difficult phase in cocrystal development [51].
The basic concept for cocrystal production is supramolecular chemistry [52]. Most of the APIs and coformers possess intermolecular bonding moieties, the most common of which is hydrogen bonding [53]. Various screening procedures are used to determine the likelihood of supramolecular synthon formation between the binding sites of the drug and coformer [54], the experimental product is then synthesized and characterized to identify a successful hit. Cocrystal screening has had a lot of success, from using synthon matching to complicated Insilco coformer screening techniques [55].
Cocrystal screening approaches have evolved through time and can be divided into conceptual, in-silico, and practical methods [56]. Due to a dearth of research into the kinetics of cocrystallization, no single strategy assures cocrystal formation [57]. The data gathered through theory and in-silico methods aid in justifying the experimental procedure and fine-tuning the cocrystal’s attributes [58].

3.1. Hypothetical/Theoretical Methods

The process of formulating cocrystals has been explained using several conceptual approaches such as “Hydrogen bonding propensity, Cambridge Structure Database, supramolecular synthon, pKa values, and Hansen solubility parameters”.

3.1.1. Hydrogen Bonding

The motif of hydrogen bonding synthons of cocrystals may be based on Etter’s empirical hydrogen-bonding guidelines for designing the hydrogen-bonded solids [59].
Etter investigated the predictability of hydrogen bonding in known crystal structures and presented a set of empirical hydrogen-bonding principles that regulates the H-bond assembly of chemical compounds forming crystals [60]. Etter’s three hydrogen-bonding principles are as follows:
  • “All good proton donors and acceptors are used in hydrogen bonding”
  • “Six-membered-ring intramolecular hydrogen bonds form in preference to intermolecular hydrogen bond”
  • “The remaining best proton donors and acceptors, after intramolecular hydrogen-bond formation, form intermolecular hydrogen bonds with one another”
These guidelines assist in discovering excellent donor as well as acceptor groups and enable the development of multi-coformer cocrystals. This technique was effectively used by V J Nikam and his co-authors to identify the coformers for nebivolol hydrochloride [61].

3.1.2. Supramolecular Synthon Approach

Noncovalent interactions between the drug and coformers such as hydrogen bonding, van der Waals forces, and pi-pi stacking play a vital role in the formation of cocrystals [62].
In accordance with these guidelines, Desiraju [63] suggested the model of supramolecular synthons approach “Supramolecular synthons are structural units within supermolecules which can be formed and/or assembled by known or conceivable synthetic operations involving intermolecular interactions”.
Desiraju stated that “supramolecular synthons are robust entities and can be exchanged from one hydrogen bonding assembly to another”.
The following processes are involved in designing cocrystals utilizing hydrogen bonding laws and the supramolecular synthon approach:
  • “Identification of functional groups in the API and coformer molecules”.
  • “Assessment of homo interactions in the individual pure molecules of the target compounds”.
  • “Identification of the possible functional groups which may compete with the homo interactions of pure compounds”.
  • “Assessment of the likelihood of hetero interactions”.
  • “Selection of coformers based on the assessment of homo/ hetero interactions”.
This technique was effectively used by Rajendrakumar S and co-workers in screening the coformers for deferiprone [64].
Supramolecular synthons are divided into two types:
Supramolecular Homosynthon: Functions that are the same and complimentary.
Supramolecular Heterosynthons: Functions that are diverse yet complimentary.
The theory of the homo and hetero synthon approach is illustrated in Figure 5.

3.1.3. ΔpKa Rule of Three

This rule states that “if the value of ΔpKa is less than 0, the possibility of formation of cocrystal increases while, for the ΔpKa differences greater than 3 will result in the formation of salts”. ΔpKa within the range of 0 to 3 is defined to be a grey zone where no accurate prediction is possible [65].
The ΔpKa value is estimated by using the equation given below:
Δ p K a = p K a b a s e p K a a c i d  
The proton transfer possibility is determined by the ΔpKa value obtained from the above equation.
Dos Santos et al. justified the ΔpKa rule by exploring several coformers to prepare cocrystals of cinnamic acid and succeeded in improving the solubility of the low soluble trans-cinnamic acid [66]. Although the ΔpKa rule can be used for preliminary screening, it is unlikely to be sufficient for cocrystal screening.

3.1.4. Hansen Solubility Parameter

Another theoretical strategy for screening cocrystals is the Hansen solubility parameter (HSP). According to HSP, “cocrystals are considered to be miscible systems, and the miscibility of API and coformer is supposed to be related to the solubility parameter (δ)”.
As a result, estimating the solubility parameter can give a good idea of cocrystallization possibilities. The solubility parameter can be determined using the formula:
δ t = δ 2 d + δ 2 p + δ 2 h
where,
  • (δd) = Dispersion.
  • (δp) = Polarity.
  • (δh) = Hydrogen bonding.
In the literature, there are various examples of HSP being used as a cocrystal screening method [67,68].

3.2. In-Silico Techniques

Several in-silico approaches for screening cocrystals are now available and it has been possible due to better technologies and computer modelling techniques. There are a variety of strategies that may be utilized to properly forecast the production of cocrystals, which saves time and money in research and development.

3.2.1. Cambridge Structural Database (CSD)

The CSD is an important resource for determining intermolecular hydrogen bonding in crystals [69]. CSD is significant for the development in the field of chemical sciences, material sciences, and life sciences, as well as drug innovation and development. With roughly 40,000 new structures added each year, the CSD is a well-designed and efficient software [70].
CSD can also be used to forecast stable H-bond motifs, to maintain the strongest motifs throughout a set of core structures [71,72]. On the other hand, CSD also facilitates hydrogen bonding propensity (HBP), a tool for designing cocrystals [73].
To understand the behavior of molecules and intermolecular forces better within a crystal, CSD allows for the retrieval, visualization, and analysis of experimentally obtained crystallographic data. This provides details on the type of molecular interactions, geometrical alternatives, directional properties, and supramolecular synthon types involved [74].
According to Fabian’s method [75], CSD provides the most reliable cocrystal forming pairs where certain parameters are considered for every molecule, known as molecule descriptors, for example, polarity, molecular shape, size, dipole moment, a single atom, surface area, etc. The system database figures out the pairs of molecules which could develop pharmaceutical cocrystals on account of calculated molecular properties. The strongest and most effective descriptor correlation was associated with the polarity and shape of the coformers [76].
Statistical analysis using cocrystal datasets on CSD enables research groups to use virtual screening approaches to locate acceptable cocrystal forming combinations, further enabling cocrystals to be built using molecular modelling, saving both time and money [77].

3.2.2. Conductor-Like Screening Model for Real Solvents (COSMO-RS)

COSMO-RS is an alternative method that is growing in popularity as a high-performance virtual screening tool. This theory is a unique blend of quantum chemistry and solution-based thermodynamics. The difference in the residual enthalpies of both the individual components as well as the cocrystal is used to forecast the creation of the cocrystal phase, assuming interactions within the solid phase are comparable to those in the supercooled phase [78]. COSMO-RS is a high-throughput screening method that may be used for screening hydrates and assist the selection of optimal cocrystallization solvents [79]. Di Wu et al., explored the prospects of COSMO-RS with other models to calculate the cocrystal consequences of the host, 2-amino-4,6-dimethoxypyrimidine (MOP), with another 63 molecules and concluded that COSMO-RS was the most efficient and reliable technique with a total hit rate of 84.1% [80].

3.2.3. Molecular Electrostatic Potential Surface (MEPS)

The MEPS technique, which is closely similar to the COSMO-RS approach, focuses on theoretical concepts and experimental findings of intermolecular interactions in the gaseous phase rather than just the liquid phase employed in the COSMO-RS strategy. The MEPS technique considers Gibbs free energy, and while enthalpy and entropy are important for characterizing intermolecular interactions, determining them is a tough process since a tiny change in the system may have a large influence on the system’s enthalpy and entropy, even though the shift in the free energy is very minimal. Free energy is regarded as a significant parameter to measure the intermolecular interactions [81]. Barbas et al., evaluated sildenafil and resorcinol using the MEPS approach for the successful formation of their cocrystals [82]. In another study Dezhi Yang et al., successfully determined the changes in the intermolecular interaction sites of three flavanols—explicitly, kaempferol, quercetin, and myricetin—with the cocrystal conformer praziquantel using MEPS and prospered in predicting the conformational changes occurring during the formation of cocrystals [83].

3.2.4. Lattice Energy Assessment

Another virtual screening approach that considers thermodynamic principles to forecast probable coformers is lattice energy [84]. The cocrystal phase is likely to be preferred when it is thermodynamically stable than the individual elements, i.e., the cocrystals’ lattice energy should be greater than the individual elements [85].
Each molecule’s lattice energy is calculated using Equation (3) as the sum of:
E l a t t = U i n t e r + Δ E i n t r a  
where,
  • Elatt = Lattice energy.
  • Uinter = intermolecular lattice energy.
  • ΔEintra = conformational intramolecular energy.
Greater the difference between the lattice energy of cocrystal and pure elements, the higher the probabilities of cocrystallization [86]. This method is not just restricted to neutral molecules but may be used to analyze salts, solvates, and other solid and liquid systems. Despite previous methods, this method makes no assumptions regarding hydrogen bonding [87].

3.3. Pick and Trial Approaches

Cocrystallization studies with small quantities of API and coformer are used for this experimental approach. Broadly, mechanochemical synthesis methods, for example neat grinding and solvent-based grinding, are desired for screening coformers for cocrystals [88]. These approaches utilize kinetic energy as a tool for cocrystallization. Equimolar proportions containing API as well as coformer are co-grounded [89] for a predetermined length of time either using a ball mill or a mortar pestle [90].
Liquid-assisted grinding (LAG) uses a little quantity of any organic solvent to speed up the cocrystallization process [91]. Liquid-assisted grinding is often chosen over neat grinding due to its inherent benefits and a better success rate [92].
Several solution-based screening techniques, including slurry and solution crystallization, are also used in addition to mechanochemical procedures [93].
The techniques described above have their own set of restrictions that are tough to overcome. Mechanochemical techniques may sometimes fail in developing cocrystals. The problem of finding a suitable solvent for cocrystallizing in a solution-based technique is a hurdle to overcome [94]. Additionally, it is impossible to exclude the occurrence of solvate/hydrates. Another noteworthy technique for cocrystallization is a cocktail method, where more than two coformers will be co-grounded with API, saving time and money [50,95].

3.4. Thermal Screening Techniques

Several thermal screening techniques, such as differential scanning calorimetry (DSC) [96], and Hot Stage Microscopy (HSM) [97], provide a quick, time- and money-saving technique for selecting suitable coformers to prepare cocrystals.
Park et al., utilized differential scanning calorimetry (DSC), a thermal screening tool to verify the possibility of cocrystal formation between emodin (EM) with various coformers. Physical mixtures of emodin and various coformers (1:1 equimolar ratio) were prepared by mixing in glass vials for 1 min. Samples showed DSC profiles of exothermic peaks after an endothermic peak. If cocrystallization is not feasible, an endothermic peak equivalent to eutectic melting is detected, followed by separate compound melting peaks [98].
Zhou et al., investigated a cocrystal screening method using HSM. Based on theoretical assumptions, Berry et al., developed a method using HSM to witness cocrystal formation [99]. Succinic acid, a higher melting point compound, and benzamide, a lower melting point compound, formed a definite mixing zone, which may signify a new cocrystal formation. On a glass slide, the API and coformer are heated until they melt and are covered with a coverslip. After cooling, the mixture is examined under HSM to observe whether cocrystals are formed. In the case of cocrystallization, a mixing zone will appear, where the novel cocrystal phase would be evident and distinguishable from the pure components [100].
Combining the findings of DSC and HSM, these techniques may improve the effectiveness of thermal screening [101].

4. Methods for Preparing Cocrystals

Solution crystallization and solid-state crystallization are the two main types of cocrystal preparation techniques. The preparation of pharmaceutical cocrystals has been discussed by Sheetal et al. [34], and Dennis et al. [94], they further discussed in detail their different possible techniques for the preparation of efficient cocrystals.
The techniques for cocrystal preparation can be generally categorized as green synthesis techniques (also called non-solvent based), solvent-based, and supercritical fluid approaches [102]. Figure 6 includes a list of some conventional cocrystal preparation methods.

4.1. Green Synthesis Techniques or Non-Solvent Techniques

Non-solvent is also a green technique since these techniques require less or no solvent for the formation of cocrystals. They are divided into four types: LAG, extrusion, simple/neat grinding, and hot-melt extrusion [103].
The neat grinding technique involves mixing two or multiple coformers in a predetermined stoichiometry and grinding them either physically or mechanically for a certain amount of time [104].
The LAG method follows a similar approach as in neat grinding, with the exception that only a small quantity of organic solvent is used during the grinding process [105].
The extrusion process along with a single screw or twin screw helps in mixing coformers with API and is operated below the melting point of the starting material [106].
Hot-melt extrusion, on the other hand, is a specialized process that involves melting and combining coformers simultaneously through a hot screw extruder. Extrusion provides intense mixing, shear, and compact material packing (avoids the use of any solvent) that increases surface contact with coformer blends, resulting in cocrystal formation [107].

4.2. Solvent-Based Techniques

For preparing cocrystals, a range of solution-based processes is known, including evaporative, spray drying, solvent evaporation, slurry, and reactive cocrystallization [93]. A solution- or evaporative-based cocrystallization technique utilizes solvent as a medium that uses undersaturated solutions of both coformer and API [108]. This approach is well-known for producing single crystal cocrystals for diffraction research and crystal structure elucidation.
Cocrystallization via assisted evaporation is comparable to the solvent evaporation technique, where evaporation takes place at higher temperatures and/or at lower pressure [109].
The spray drying process is commonly employed to make ASDs, however new research has shown that it may also be utilized to prepare cocrystals. It is advantageous since the process is continuous, highly controlled, and rapid. Spray drying is a single-step process for converting liquids such as solutions, suspensions, or slurries to solid particles by dispersing an unsaturated liquid of both drug and coformers via a nozzle using nitrogen, where the solvent is quickly removed [110].
The slurry conversion technique (also known as isothermal slurry conversion) entails introducing solid coformers in a predetermined stoichiometric ratio to a solvent (or a solvent combination) for equilibration. This process does not involve a clear starting solution, unlike evaporative cocrystallization. Reactive cocrystallization involves mixing of coformer and API individually and later adding into either of the clear solutions. Under ambient circumstances, this causes a quick spontaneous cocrystallization [111].
In the ultrasound-assisted solution approach, suitable coformers and API are dissolved in a solvent to produce nanocrystal formulations. To avoid fragmentation and deterioration, the solution is placed in a sonicator and maintained at a constant temperature. The solution is maintained overnight to allow evaporation of the solvent and the development of cocrystals [65].

4.3. Supercritical Fluid Approach

The supercritical fluid technique involves heating a mixture of API and co-former to a temperature close to the melting point of one component (in this case, the co-former) and sustaining these conditions long enough until a single crystal from the multicomponent system is obtained. Unlike organic solvents, supercritical fluid allows relatively faster kinetics and helps in easily removing the organic solvents from the products. This technique also reduces the volume of equipment and expenses of subsequent solvent separation processes during the manufacturing of cocrystals [112].
Cocrystallization through the supercritical solvent approach requires the slurry form of both API and coformer suspended in supercritical CO2. This method utilizes only the supercritical CO2 rather than any hazardous organic solvents. Firstly, CO2 thermodynamics are tweaked to stabilize the density and solvent power. As a result, controlling the cocrystallization process between the cocrystal components shall be possible. The use of this approach resulted in a higher rate of cocrystallization. This is caused by a large amount of mass transfer via convection, which is aided by mixing cocrystal components in a CO2 slurry, resulting in entire cocrystallization and also a pure cocrystal [113].
The Rapid Expansion of Supercritical Solvents (RESS) technique involves saturating the supercritical fluid (supercritical CO2) with an API and a co-former before depressurizing the CO2 phase through a nozzle into a drying chamber with atmospheric pressure. Unfortunately, this technique demands API and co-former solubility in supercritical CO2, which most pharmaceutical compounds fail to have [114].
The Supercritical Antisolvent Cocrystallization (SAC) process utilizes supercritical CO2 as an antisolvent. In contrast to RESS, this technique requires less soluble API and coformer throughout supercritical CO2 in order to precipitate out the two molecules as one. Once the components are added to the vessel, CO2 dissolves in the solvent, causing a concurrent expansion of the volume and lowering the solvent’s solubility, leading to precipitation [115].
Batch gas antisolvent (BGAS) is a process in which a solution containing both API and coformer is saturated in a vessel with carbon dioxide elevated with pressure until cocrystallization completes. Another method is a semi-continuous supercritical antisolvent (SSAS) technique, in which an API-coformer solution is pumped through a nozzle into a highly pressured vessel containing supercritical CO2 [30].

5. Characterization of Cocrystals

Specific to cocrystal analysis, microscopic, spectroscopic, and thermal methods are widely utilized. To assess developability into a marketable dosage form, physicochemical parameters such as crystallinity, melting temperature, stability, dissolution, and solubility are examined in the same way as any other solid form [116]. The “melting point is a fundamental physical property, which is a thermodynamic process determined by the temperature of equilibrium between the solid and liquid phase when the free energy of transition is equal to zero” [117]. Differential scanning calorimetry (DSC) is a quick way to acquire a melting point and thermal data such as melting enthalpy [118]. Differential Thermal Analysis (DTA) and Thermogravimetry (TG), in addition to the DSC, have been used to determine various factors such as polymorphisms, glass transitions, hydration, decomposition, and stability. PXRD and SCXRD (powder X-ray diffraction and single-crystal X-ray diffraction, respectively) offer structural information, crystallinity degree, and crystal size [119]. SCXRD has always seemed to be the gold standard, providing complete 3D structural information as well as validating composition, packing, and hydrogen bond configurations [120]. Other spectrophotometric methods, such as infrared (IR), Raman, and nuclear magnetic resonance (NMR), provide extra information that can be used to investigate noncovalent interactions such as hydrogen bonds and associated supramolecular synthons, as well as reflect differences in intermolecular interactions among several crystal structures [121].
Surface topography, particle size, and optical characteristics may all be studied using scanning electron microscopy (SEM) and optical microscopy [122]. Hot-stage microscopy (HSM) integrates microscopy and thermal analysis to offer solid-state characterization as a tool for temperature, allowing for the detection of localized thermal non-uniformity on the sample surface that is overlooked by DSC but visible in HSM pictures [123]. When it comes to orally administered dosage forms, however, stability in both the solid and liquid states is critical. Thermal and relative humidity stress (RHS) was used to evaluate possible solid-state alterations. Thermal methods and PXRD are widely employed in conjunction with high-performance liquid chromatography (HPLC) for degradation testing [124].
Although the characterization techniques for several oral dosage forms serve to be similar to that of the cocrystals, this makes it difficult to differentiate between the salt and a cocrystal. The proton transfer can be in principle estimated by solid-state NMR and/or IR spectra. However, the interpretation of the results in a particular system may be tricky. Still, the SCXRD is an ultimate method for salt/cocrystal determination, while the others are indirect and are easy to misinterpret.

6. Anti-Cancer Drug Cocrystals

Teva Pharmaceuticals recently created a cocrystal of anticancer drug ibrutinib with fumaric acid, which is prescribed for chronic lymphocytic leukaemia. They demonstrated that it has similar solubility to the original medication but is more stable. There is plenty of room for numerous additional cocrystals to develop a pathway. Many researchers have designed several numbers of cocrystals to date to enhance the physicochemical parameters of several anticancer drugs and succeeded in designing the same. Several cocrystals of anticancer drugs are reported in Table 1. Due to their pKa similarities, carboxylic acids, polyphenols, carboxyl amides, and amides are frequently found as co-formers in anticancer drug cocrystals. Most anticancer drugs are either weakly acidic or weakly basic and a few anticancer drugs such as paclitaxel are zwitterionic, too [125].
A pKa range between 0 to 3 of drug and coformer aids in the formation of cocrystals, as stated in the pKa rule of three. da Silva et al. prepared cocrystals of 5-Fluorocytosine based on the ΔpKa rule and achieved several cocrystals with adipic acid, benzoic acid, succinic acid, terephthalic acid, and malic acid [126].
Nicotinamide and its derivatives are the most commonly used co-formers with anticancer drugs. Since both the amide and pyridine groups are present in the compounds, they can interact as primary synthons.
Hongyan et al. worked on resveratrol cocrystals with nicotinamide and isonicotinamide as coformers and found a huge difference in the drug release patterns and crystal patterns, though the only difference in both the coformers is the placement of the nitrogen atom in the pyridine ring. The only justification for this can be the difference in the gas phase and synthon binding energies between both cocrystals formed [127].
Chenxin et al., investigated two different coformers of palbociclib with resorcinol and orcinol. Cocrystals of palbociclib–resorcinol displayed a better powder dissolution than that of the pure palbociclib. Cocrystals of palbociclib not only improved the dissolution of the drug but also improved the bioavailability which was the greatest drawback of Palbociclib [128].
Adam et al., synthesized cocrystals of quercetin an antineoplastic drug with several coformers such as caffeine, isonicotinamide, and theobromine. From the obtained dissolution profiles, they proved that cocrystals of quercetin–caffeine had improved the solubility up to 14-fold when compared to that of pure quercetin. They also performed kinetic studies which proved the improved absorption of the drug through cocrystal technique rather than pure drug where the area under the curve (AUC) was almost 10-fold greater than with the pure quercetin [129].
Urooj et al. prepared cocrystals of nandrolone with salicylic acid by using a grinding technique. They have evaluated the anticancer activity of the cocrystals and the pure drug against HeLa cell lines. The results obtained from the cell line studies proved that cocrystals of nandrolone–salicylic acid. Researchers also found the cocrystals had a synergistic effect and the cocrystals were found to be a potent anticancer agent with anticancer activity comparable to the standard drugs [130].
Yue-Ming et al. utilized the advantages of cocrystals and prepared a recombinant strategy with cocrystallization and nano micelle formation of 5-Fluorouracil with L-Proline as a coformer. The investigators proved that the solubility of the cocrystals was much higher i.e., up to 4.60-fold, and permeability up to 3.89-fold when compared to the pure drug. Eventually, cocrystals also provided excellent anti-tumor activities and restricted their proliferation. Additionally, the prominent slow release of the drug helped in providing a sustained release, which in turn helped to develop a formulation at a lower dose with higher efficacy [131].
Apatinib has a major issue of solubility and hence low oral bioavailability. To overcome this issue Bin et al., reported three innovative cocrystals of apatinib with coformers sebacic acid, adipic acid, and mandelic acid and tried to improve the solubility of the API using the cocrystallization technique. From the studies, they proved that apatinib–sebacic acid cocrystal was successful in increasing the solubility of apatinib up to 7.5-fold. Though solubility has increased to a greater extent, kinetic study results had shown a slight decline in the bioavailability of the apatinib salt form. The stability reports reveal that cocrystals did not convert back to the original apatinib form; this suggests that apatinib–sebacic acid cocrystals have an aqueous stability advantage over the original form [132].
The above-mentioned article proves that alteration of solubility is not just a single parameter that contributes to the bioavailability enhancement of cocrystals. Despite the fact that solubility is a limiting factor in drug oral bioavailability, it is clear that it is not the sole determinant. It is probable that altered metabolism has a significant impact on oral bioavailability.
In general terms, most of the researchers worked on multiple anticancer drugs for enhancing their solubility and stability using various cocrystallization techniques.
Yueming et al., worked on BCS class III drug tegafur to enhance the permeability and its antitumor activity. Researchers have used syringic acid as a coformer to prepare the tegafur cocrystals. The present tegafur–syringic acid cocrystals synchronously enhanced both solubility and permeability. The in vivo pharmacokinetic studies reported a prolonged half-life and enhanced bioavailability when compared to the pure drug. Tegafur–syringic cocrystals were tested on the tumor cells along with the pure drug and displayed a synergistic anti-tumor effect [133].
Meiqi Li and colleagues worked on emodin, a natural dye with various pharmacological activities, such as anticancer, antioxidative activities, anti-inflammatory, and cathartic. Emodin color was changed from yellow to brighter red through cocrystallization using mainly nicotinamide (a colorless coformer) and improving its optical properties [134].
Occasionally, cocrystallization of a drug may not constantly result in solubility enhancement, but instead, it could increase the dissolution rate. Faster dissolution of cocrystals reduces the time to absorb drug from the gastrointestinal tract (decreased Tmax) followed by phase transformation to a less soluble form.
Shiraki et al. prepared exemestane cocrystals with coformer maleic acid and achieved excellent solubility enhancement and faster dissolution. Though solubility was improved, faster dissolution led to transformation into the unstable crystalline form [135].

Drug–Drug Cocrystals

“Drug–drug cocrystals are composed of two or more APIs established into a new class of solid forms of APIs”. Drug–drug cocrystals are the recent trending approach and are attracting more attention due to their low-cost experimentation and low risk and enhance the physicochemical and biopharmaceutical properties of both the drugs [136]. In order to address the rising need for more effective combination medications such as drug–drug cocrystals, conserving the individual physicochemical characteristics of an API is crucial in such multidrug treatments [137].
Xuan et al. successfully achieved an extended release by preparing drug–drug cocrystals of isoniazid with curcumin, which proved that cocrystals can stand as a promising strategy for achieving an extended release without the use of any polymers [138]. Drug–drug cocrystals not only enhance the properties of any of the drugs present but also are a synergistic pathway during the treatment cycle. When compared to the other dual drug therapy, drug–drug cocrystals may enhance physicochemical attributes such as dissolution, bioavailability, improved solid-state stability, half-life, etc. Even after having several benefits such as dual drug therapy, drug–drug cocrystals have been very rarely designed in the pharmaceutical field due to their own disadvantages. Wang et al. reviewed the future prospects of several drug–drug cocrystals and mentioned multiple factors helping in cocrystal formation. The synthon approach has always been the primary go-to for selecting suitable coformers, but drug–drug cocrystals require a suitable drug with suitable properties such as solubility, compatibility, and less toxicity with minimum adverse effects, which makes selection a quite challenging process. In addition to this, combinations for creating drug–drug cocrystals are also limited since they are mostly reliant on the properties of the drug than the conventional synthon approach [139].
Some of the drug–drug cocrystals available for anticancer drugs during this decade are listed below in Table 2.
Yin and his co-workers worked with lobaplatin and multiple flavonoids and prospered in obtaining the drug–drug cocrystals. Though they couldn’t achieve any enhanced solubility via the cocrystals they were able to crack down the stability issues of the parent drug by decreasing the hydrolysis effect on the parent molecule. On the other hand, some cocrystals failed to improve the solubility of the molecules, rather, they provided sustained release rates which helped to achieve added antitumor activity [140].
da Silva and her co-workers worked on 5-Fluorocytosine with 5-Fluorouracil drug–drug cocrystals. They created drug–drug cocrystals using solvent drop grinding, a green synthesis technique and achieved purely stable cocrystals which are a very important step in cocrystal development [141].
Shridhar et al. cocrystallized gefitinib with furosemide and succeeded in forming cocrystals using solution crystallization. Cocrystal structure study discovered that the components in the crystal lattice are tightly held by strong hydrogen bonds leading to compact crystal packing and enhanced stability. As mentioned, the stronger hydrogen bonds lead to form a more insoluble form of gefitinib than its other stable polymorphs. This increased the retention time of the drug at the place of action [142].
Liu et al. carried out cocrystallization between dihydromyricetin with pentoxifylline and estimated that the solubility of pentoxifylline in an aqueous medium shrinks significantly to 100-fold, while the solubility of dihydromyricetin increases slightly by 1.2-fold. This helps in maintaining a sustained release of both the drugs for up to twelve hours, maintaining the saturation concentration. This drug–drug cocrystal mechanism helped in maintaining a synergetic anticancer effect on the cancer cells [143].
Yin et al. investigated drug–drug cocrystals of oxaliplatin using flavonoids such as baicalein and naringenin and were successful in achieving drug–drug cocrystals with a slow-release rate. Researchers also observed a delayed rate of hydrolysis, which further increases the stability of oxaliplatin. The achieved drug–drug cocrystals of oxaliplatin with baicalein have shown a substantial effect by inhibiting the cancer cells [144].
Jie Wang and his coworkers investigated on a new equimolar drug–drug cocrystal of temozolomide with hesperetin. The formed cocrystals not only improved the stability of the individual drugs but also established an extended dissolution rate which directly helps in the improvement of the physicochemical characteristics of the drug [145].
As listed below, the work on drug–drug cocrystals to date is limited; only four of them were marketed and a few are still in the pipeline for FDA approvals. The development of drug–drug cocrystals has several drawbacks. There is a need for pharmaceutical researchers to develop robust techniques for forecasting drug–drug cocrystals formation. As an outcome, it may fetch a novel prospect for drug–drug cocrystals and new applications to the conventional drug molecules.
Table 1. A list of reported cocrystals of anti-cancer drugs, cocrystal preparation techniques, and its improved parameters from pure API.
Table 1. A list of reported cocrystals of anti-cancer drugs, cocrystal preparation techniques, and its improved parameters from pure API.
Anti-Cancer DrugsCoformers UsedPreparation TechniquesEnhanced
Parameters from Pure Drug
Reference
4,4′-ethylenebispyridineFlavonoidsSolvent Evaporation methodImproved dissolution behaviour, and anti-tumor activity[146]
5-FluorouracilFerulic acidSolvent assisted co-grindingImproved solubility and permeability[147]
5-FluorouracilNicotinamide and Iso-NicotinamideSolvent-free grinding techniqueImproved solubility and stability[148]
5-FluorouracilProlineSolvent evaporation technique, and Liquid assisted grindingImproved solubility, permeability[131]
5-FluorouracilSuccinic acid, Malic acid Cinnamic acid, and Benzoic acidNeat grinding and slow solvent evaporationImproved anticancer activity[149]
5-FluorouracilUrea,
Thiourea, Pyrazinamide
Supercritical solvent techniqueImproved solubility[150]
5-FluorouracilUrea,
Thiourea and Acetanilide
Solid-state grinding methodImproved solubility[151]
5-FluorouracilGentisic acid, 3,4-Dihydroxybenzoic, 4-AminopyridineSolvent-assisted grinding and solution crystallizationImproved permeability and antitumor efficacy[152]
6-Mercaptopurine monohydrate2,4-Dihydroxybenzoic acid and Hydroxybenzoic acidSlurry reactive crystallizationImproved solubility and stability[153]
6-Mercaptopurine monohydrateZinc Tri-fluoro methane sulfonateSolvent evaporationImproved solubility and dissolution[154]
ArtemisininAcetylene dicarboxylic acidGreen synthesis techniqueImproved solubility and dissolution[155]
AxitinibGlutaric acidSolvent evaporation techniqueImproved solubility and stability[156]
AxitinibFumaric acid, Cinnamic acidLiquid assisted grinding and slurry methodsImproved solubility, stability, and oral bioavailability[157]
BaicaleinNicotinamideSlurry methodImproved solubility, dissolution, and bioavailability[158]
Betulinic acidAscorbic acidSlow solvent evaporationImproved bioavailability, solubility[159]
CeritinibNicotinamide, QuercitinNeat grindingImproved solubility and stability[160]
CoumarinThioureaNeat grinding methodImproved antioxidant activity and solubility[161]
CurcuminN-acetylcysteineAnti-Solvent gas
Supercritical fluid technique
Improved oral bioavailability[162]
CurcuminAscorbic acidSolvent evaporation methodImproved solubility, stability, and biological activities[163]
CurcuminTrimesic acidSolid-state grinding, slow cooling crystallization, Evaporative crystallizationEnhanced dissolution[164]
CurcuminN-acetylcysteineSupercritical fluid technologyImproved solubility and dissolution[165]
CytarabineNicotinamideSlow solvent evaporation methodImproved solubility[166]
DabrafenibFumaric acid, Succinic acid and Adipic acidSolvent drop grinding methodImproved solubility and dissolution[167]
DasatinibN-methyl-4-hydroxybenzoate,
Nicotinamide, Ethyl gallate, Methyl gallate, Propyl gallate, Vanillin
Solvent evaporation techniqueImproved solubility, stability, and oral bioavailability[168]
DocetaxelNicotinamideSlow solvent evaporation techniqueImproved solubility, dissolution and penetration[169]
DoxorubicinMaravirocSlow solvent evaporation methodImproved solubility[166]
EmodinNicotinamideSolvent evaporationImproved solubility, oral absorption, and bioavailability[170]
GefitinibIsonicotinamide and
Vanillin
Solvent evaporation methodImproved intrinsic dissolution, bioavailability, and cell inhibition response[171]
IbrutinibHydroxy benzoic acid, Succinic acid, Hydroxy naphthoic acidSlurry method, ultrasonic methodImproved solubility, stability[65]
ImatinibSyringic acidSlurry methodImproved solubility, dissolution, and permeation[172]
LapachoneResorcinolMechanical grindingImproved solubility and oral bioavailability[173]
LenalidomideUrea and 3,5-dihydroxybenzoic acidSolvent evaporation techniqueImproved solubility[174]
LenalidomideGallic acidLiquid assisted grindingImproved solubility, stability, and dissolution[175]
LenvatinibSulfamerazineSlow solvent evaporation methodImproved Solubility and stability[176]
LenvatinibSalicylic acidSlow solvent evaporation methodImproved Solubility and stability[176]
LenvatinibSulfamerazine and salicylic acidSlurry reactive crystallizationImproved solubility, stability, and dissolution[176]
LuteolinCaffeineLiquid assisted grinding and rapid solvent removal methodImproved solubility, thermal stability, and bioavailability[177]
Megestrol acetateSaccharinSlow solvent evaporation techniqueImproved dissolution[178]
NandroloneSalicylic acidGrinding methodImproved Anticancer activity[130]
Nandrolone3-amino-1,2,4-triazoleSolution reflux methodImproved Anticancer activity[130]
PalbociclibResorcinolSolvent evaporation techniqueImproved solubility, bioavailability, and biosafety[128]
PalbociclibOrcinolSolvent evaporation techniqueImproved solubility and dissolution[128]
PterostilbenePicolinic acidLiquid assisted grindingImproved oral bioavailability[179]
RegorafenibMalonic acid, Glutaric acid and Pimelic acidLiquid assisted grindingImproved solubility and dissolution[180]
ResveratrolNicotinamideSupercritical CO2 by the gas antisolvent (GAS) techniqueImproved dissolution and bioavailability[181]
TegafurSyringic acidLiquid assisted grinding and solvent evaporationProlonged half-life and increased bioavailability[133]
TemozolomideSuccinic acid, Oxalic acidSolvent evaporation techniqueImproved solid-state stability, bioavailability, dissolution[182]
Zoledronic acidTartaric acid and NicotinamideSlurry method, solvent evaporation, and dry grindingImproved solubility[183]
Table 2. Drug–drug cocrystals of anti-cancer drugs reported in the last decade.
Table 2. Drug–drug cocrystals of anti-cancer drugs reported in the last decade.
Drug
Combinations
Therapeutic
Category
Preparation MethodsObservationsReference
Dihydromyricetin
with Pentoxifylline
Anticancer and Anti ischemicSlurry methodImproved solubility and synergistic anticancer effect[143]
Lobaplatin
with
Quercetin, Myricetin, Fisetin, Naringenin, and Luteolin
Anticancer and FlavonoidsSlow solvent evaporation methodImproved dissolution, stability, and antitumor effect[140]
Oxaliplatin with BaicaleinAnticancer and FlavonoidsSolvent evaporation methodReduced release rate[144]
Oxaliplatin with NaringeninAnticancer and FlavonoidsSolvent evaporation methodDelayed hydrolysis[144]
Temozolomide with HesperetinAnticancer and FlavonoidsLiquid assisted grindingImproved stability and extended dissolution[145]
Theophylline with QuercetinAnticancer drugsLiquid assisted grindingImproved thermal properties, stability, solubility, and anticancer effect[184]
5-fluorouracil with SorafenibAnticancer drugsSlurry methodImproved solubility, stability, and antitumor effect[185]
5-fluorouracil with RegorafenibAnticancer drugsHetero seeding methodImproved solubility, stability, and antitumor effect[185]
Propylthiouracil with kaempferolAnti-thyroid and Anticancer drugsSolvent evaporationImproved bioavailability[186]
Erlotinib with FurosemideAnticancer and Diuretic drugsLiquid assisted grindingImproved solubility, thermal stability, and dissolution[187]
Gefitinib with MefenamicAnticancer and NSAIDsLiquid assisted grindingImproved solubility[187]
Oxaliplatin with BaicaleinAnticancer and FlavonoidsSolution crystallization and liquid assisted grindingImproved physicochemical properties[144]
Oxaliplatin with NaringeninAnticancer and FlavonoidsSolution crystallizationImproved physicochemical properties[144]
5-fluorouracil with
kaempferol
Anticancer and FlavonoidsSlurry methodImproved stability and dissolution[188]
Berberine chloride with
Myricetin and Dihydromyricetin
Anticancer and FlavonoidsGreen synthesisEnhanced solubility and synergistic anticancer effect[136]
Luteolin with IsoniazidAnticancer and AntituberculosisLiquid assisted grinding and rapid solvent removal methodImproved solubility, thermal stability, and bioavailability[177]
Doxorubicin HCl with Quercetin hydrate and NaringinAnticancer and FlavonoidsSlow solvent evaporation techniqueImproved solubility[189]
5-Fluorouracil with AspirinAnticancer and NSAIDsSolid-state grinding methodImproved solubility[151]
Emodin with Berberine ChlorideAnticancer drugsSlurry methodImproved sustained-release hence absorption[190]
Imatinib mesylate with 5-fluorouracilAnticancer drugsSolvent evaporation techniqueImproved solubility and stability[191]
Dacarbazine with QuercetinAnticancer drugsSolvent drop grindingImproved solubility and dissolution[192]
Temozolomide with HesperitinAnticancer and FlavonoidsLiquid assisted grindingImproved solubility, stability, and dissolution[145]

7. Prospects of Cocrystals in the Future

Anticancer drugs have consistent growth in the market and the rise in cancer incidents year to year increases the demand for these drug molecules. This makes those drugs a suitable aim for cocrystallization and overcoming their solubility issues, as discussed previously. Furthermore, it allows scientists to file patents for updated versions of these APIs.
Even though after having all these practical advantages only a couple of anticancer drug cocrystals may hit the market and a few more are still at the clinical trial stage, ibrutinib–fumaric acid cocrystals, used as an anticancer drug to treat chronic lymphocytic leukaemia, are under the process of FDA approval [193].
In this review, several numbers of anticancer drug cocrystals and drug–drug cocrystals were already addressed but the current market situation is contradictory to their achievements.
In the field of cocrystal synthesis, formulation approaches are always evolving, with several procedures to attain the same goal and the search for more dependable methods that require very little time, solvent, and other resources. Earlier, cocrystals were not much explored due to the absence of current technologies, because of which the concept of cocrystals was not well understood, resulting in a challenging scale-up procedure. Researchers from the pharmaceutical industries have found some significance in developing technologies for cocrystal preparation. Nowadays in-silico and several virtual methods are being adapted for coformer selection and cocrystal design, which helps in reducing practical time and chemical consumption and bringing out a new standpoint in drug development without modifying the molecules chemically [194]. Drug–drug cocrystals are an innovative technique for overcoming some of the issues that typical fixed-dose combinations have. Ashwini et al. reviewed many such case studies and proved numerous abilities of cocrystals; several attempts of cocrystallization of either two or more APIs have demonstrated substantial benefits by enhancing drug efficacy, synergic effects, and mainly the physicochemical parameters without changing API chemically. This leads us to conclude that there is a lot more to explore and utilize the extended-release capabilities of cocrystals, thus proving that there is still a lot of room for advancement in this discipline, allowing for more control over disease treatment, side effects, and dose load [195].
As reported by Zaworotko et al., pharmaceutical cocrystals can be a torch bearer for several such molecules, which require an eye on their drug release patterns to improve their bioavailability. Cocrystals can lead to a new path for all those APIs impacted with poor physicochemical properties directly restricting the solubility and further their oral bioavailability [37].
Nowadays, versatile formulations are capable in imaging, and theranostics have achieved more limelight and been a research interest for many. Wang et al. developed cocrystals that were capable of imaging using near-IR photothermal cocrystals. These cocrystals were synthesized by molecular self-assembly with two small molecules helpful in photothermal imaging. Photothermal imaging has potential applications in various fields which include photothermal therapy (PTT) photothermal/photoacoustic (PT/PA) imaging, photo-thermal-electric devices, and shape-memory devices which are generally utilized in anticancer therapy [196].

8. Conclusions

The steady progress in the total number of research publications on the subject of pharmaceutical cocrystals, as mentioned earlier, indicates the prominence of cocrystallization in the medical field and also shows the keen interest in the researchers. From this review, it seems to be very clear that pharmaceutical cocrystals still carry the limelight around them for their adaptability without changing the scaffold of the chemical structure. However, there are still certain obstacles to overcome, such as coformer selection, physicochemical characterization, and formulation. Successful cocrystal development may be achieved with appropriate API–coformer identification and formulation design. The current review presented different cocrystal screening techniques, including the green screening techniques, which are more advantageous. Several benefits of cocrystallization and fundamentals have been discussed. Nevertheless, as time has passed and the field has advanced, researchers in this area have established several novel techniques, progressively effortless to facilitate cocrystallization procedures, to conquer their preceding limitations. In summary, this review covered several cocrystals and drug–drug cocrystals of anticancer drugs and how these cocrystals have become more popular as a technique to increase the bioavailability of drugs that are water-insoluble. Additionally, the challenges to be addressed and their prospects have been discussed in detail.

Author Contributions

Conceptualization: D.D.K. and M.R.; writing—original draft preparation, D.D.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors would like to thank Manipal College of Pharmaceutical Sciences, MAHE, Manipal, India. Authors are also thankful to Biorender, an image making tool.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Novais, M.V.M.; Gomes, E.R.; Miranda, M.C.; Silva, J.O.; Gomes, D.A.; Braga, F.C.; Pádua, R.M.; Oliveira, M.C. Liposomes Co-Encapsulating Doxorubicin and Glucoevatromonoside Derivative Induce Synergic Cytotoxic Response against Breast Cancer Cell Lines. Biomed. Pharmacother. 2021, 136, 111123. [Google Scholar] [CrossRef]
  2. Layek, B.; Gidwani, B.; Tiwari, S.; Joshi, V.; Jain, V.; Vyas, A. Recent Advances in Lipid-Based Nanodrug Delivery Systems in Cancer Therapy. Curr. Pharm. Des. 2020, 26, 3218–3233. [Google Scholar] [CrossRef] [PubMed]
  3. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  4. Global Burden of Disease Cancer Collaboration. Global, Regional, and National Cancer Incidence, Mortality, Years of Life Lost, Years Lived with Disability, and Disability-Adjusted Life-Years for 32 Cancer Groups, 1990 to 2015: A Systematic Analysis for the Global Burden of Disease Study. JAMA Oncol. 2017, 3, 524–548. [Google Scholar] [CrossRef] [PubMed]
  5. Lambert, J.M.; van Delft, F.L. Chapter 1 Introduction to Antibody–Drug Conjugates. In Cytotoxic Payloads for Antibody–Drug Conjugates; Royal Society of Chemistry: London, UK, 2021; pp. 1–31. [Google Scholar] [CrossRef]
  6. Cooper, G.M. The Development and Causes of Cancer. In The Cell: A Molecular Approach, 2nd ed.; Sinauer Associates Inc.: Sunderland, MA, USA, 2000. [Google Scholar]
  7. Nair, A.; Amalraj, A.; Jacob, J.; Kunnumakkara, A.B.; Gopi, S. Non-Curcuminoids from Turmeric and Their Potential in Cancer Therapy and Anticancer Drug Delivery Formulations. Biomolecules 2019, 9, 13. [Google Scholar] [CrossRef]
  8. Bayón-Cordero, L.; Alkorta, I.; Arana, L. Application of Solid Lipid Nanoparticles to Improve the Efficiency of Anticancer Drugs. Nanomaterials 2019, 9, 474. [Google Scholar] [CrossRef]
  9. Sawicki, E.; Schellens, J.H.M.; Beijnen, J.H.; Nuijen, B. Inventory of Oral Anticancer Agents: Pharmaceutical Formulation Aspects with Focus on the Solid Dispersion Technique. Cancer Treat. Rev. 2016, 50, 247–263. [Google Scholar] [CrossRef]
  10. Sohail, M.F.; Rehman, M.; Sarwar, H.S.; Naveed, S.; Salman, O.; Bukhari, N.I.; Hussain, I.; Webster, T.J.; Shahnaz, G. Advancements in the Oral Delivery of Docetaxel: Challenges, Current State-of-the-Art and Future Trends. Int. J. Nanomed. 2018, 13, 3145–3161. [Google Scholar] [CrossRef]
  11. Han, W.; Xie, B.; Li, Y.; Shi, L.; Wan, J.; Chen, X.; Wang, H. Orally Deliverable Nanotherapeutics for the Synergistic Treatment of Colitis-Associated Colorectal Cancer. Theranostics 2019, 9, 7458–7473. [Google Scholar] [CrossRef]
  12. Gala, U.H.; Miller, D.A.; Williams, R.O. Harnessing the Therapeutic Potential of Anticancer Drugs through Amorphous Solid Dispersions. Biochim. Biophys. Acta (BBA) Rev. Cancer 2020, 1873, 188319. [Google Scholar] [CrossRef]
  13. Chavda, H.; Patel, C.; Anand, I. Biopharmaceutics Classification System. Syst. Rev. Pharm. 2010, 1, 62. [Google Scholar] [CrossRef]
  14. Suresh, K.; Nangia, A. Curcumin: Pharmaceutical Solids as a Platform to Improve Solubility and Bioavailability. CrystEngComm 2018, 20, 3277–3296. [Google Scholar] [CrossRef]
  15. Suzuki, M.; Takebe, G.; Takagi, T.; Tsukada, H. Characterization of Novel Paclitaxel Nanoparticles Prepared by Laser Irradiation. Chem. Pharm. Bull. 2022, 70, 269–276. [Google Scholar] [CrossRef]
  16. Parvataneni, D.M.; Devraj, R.; Mangamoori, L.N. Micelles Entrapped Microparticles Technology: A Novel Approach to Resolve Dissolution and Bioavailability Problems of Poorly Water Soluble Drugs. J. Microencapsul. 2020, 37, 254–269. [Google Scholar] [CrossRef]
  17. Ashrafizadeh, M.; Ahmadi, Z.; Mohamadi, N.; Zarrabi, A.; Abasi, S.; Dehghannoudeh, G.; Tamaddondoust, R.N.; Khanbabaei, H.; Mohammadinejad, R.; Thakur, V.K. Chitosan-Based Advanced Materials for Docetaxel and Paclitaxel Delivery: Recent Advances and Future Directions in Cancer Theranostics. Int. J. Biol. Macromol. 2020, 145, 282–300. [Google Scholar] [CrossRef] [PubMed]
  18. Guo, W.; Du, S.; Lin, Y.; Lu, B.; Yang, C.; Wang, J.; Zeng, Y. Structural and Computational Insights into the Enhanced Solubility of Dipfluzine by Complexation: Salt and Salt-Cocrystal. New J. Chem. 2018, 42, 15068–15078. [Google Scholar] [CrossRef]
  19. Al-Obaidi, H.; Granger, A.; Hibbard, T.; Opesanwo, S. Pulmonary Drug Delivery of Antimicrobials and Anticancer Drugs Using Solid Dispersions. Pharmaceutics 2021, 13, 1056. [Google Scholar] [CrossRef] [PubMed]
  20. Khatun, B.; Baishya, P.; Ramteke, A.; Maji, T.K. Study of the Complexation of Structurally Modified Curcumin with Hydroxypropyl Beta Cyclodextrin and Its Effect on Anticancer Activity. New J. Chem. 2020, 44, 4887–4897. [Google Scholar] [CrossRef]
  21. Lerata, M.S.; D’Souza, S.; Sibuyi, N.R.S.; Dube, A.; Meyer, M.; Samaai, T.; Antunes, E.M.; Beukes, D.R. Encapsulation of Variabilin in Stearic Acid Solid Lipid Nanoparticles Enhances Its Anticancer Activity in Vitro. Molecules 2020, 25, 830. [Google Scholar] [CrossRef] [Green Version]
  22. Gomes, H.I.O.; Martins, C.S.M.; Prior, J.A.V. Silver Nanoparticles as Carriers of Anticancer Drugs for Efficient Target Treatment of Cancer Cells. Nanomaterials 2021, 11, 964. [Google Scholar] [CrossRef]
  23. Sathisaran, I.; Dalvi, S.V. Engineering Cocrystals of Poorly Water-Soluble Drugs to Enhance Dissolution in Aqueous Medium. Pharmaceutics 2018, 10, 108. [Google Scholar] [CrossRef] [PubMed]
  24. Alshahrouri, B.; Yang, F.; Schwing, Q.; Dürig, T.; Fassihi, R. Hot-Melt Extrusion Based Sustained Release Ibrutinib Delivery System: An Inhibitor of Bruton’s Tyrosine Kinase (BTK). Int. J. Pharm. 2021, 607, 120981. [Google Scholar] [CrossRef] [PubMed]
  25. Caruana, R.; Montalbano, F.; Zizzo, M.G.; Puleio, R.; Caldara, G.; Cicero, L.; Cassata, G.; Licciardi, M. Enhanced Anticancer Effect of Quercetin Microparticles Formulation Obtained by Spray Drying. Int. J. Food Sci. Technol. 2022, 57, 2739–2746. [Google Scholar] [CrossRef]
  26. Rizwanullah, M.; Alam, M.; Harshita; Mir, S.R.; Rizvi, M.M.A.; Amin, S. Polymer-Lipid Hybrid Nanoparticles: A Next-Generation Nanocarrier for Targeted Treatment of Solid Tumors. Curr. Pharm. Des. 2020, 26, 1206–1215. [Google Scholar] [CrossRef]
  27. Abuzar, S.M.; Hyun, S.-M.; Kim, J.-H.; Park, H.J.; Kim, M.-S.; Park, J.-S.; Hwang, S.-J. Enhancing the Solubility and Bioavailability of Poorly Water-Soluble Drugs Using Supercritical Antisolvent (SAS) Process. Int. J. Pharm. 2018, 538, 1–13. [Google Scholar] [CrossRef] [PubMed]
  28. Tran, P.H.L.; Lee, B.-J.; Tran, T.T.D. Strategies and Formulations of Freeze-Dried Tablets for Controlled Drug Delivery. Int. J. Pharm. 2021, 597, 120373. [Google Scholar] [CrossRef]
  29. Khanfar, M.; Al-Remawi, M.; Al-Akayleh, F.; Hmouze, S. Preparation and Evaluation of Co-Amorphous Formulations of Telmisartan—Amino Acids as a Potential Method for Solubility and Dissolution Enhancement. AAPS PharmSciTech 2021, 22, 112. [Google Scholar] [CrossRef]
  30. Ngilirabanga, J.B.; Samsodien, H. Pharmaceutical Co-Crystal: An Alternative Strategy for Enhanced Physicochemical Properties and Drug Synergy. Nano Sel. 2021, 2, 512–526. [Google Scholar] [CrossRef]
  31. Bharate, S.S. Recent Developments in Pharmaceutical Salts: FDA Approvals from 2015 to 2019. Drug Discov. Today 2021, 26, 384–398. [Google Scholar] [CrossRef]
  32. Gupta, D.; Bhatia, D.; Dave, V.; Sutariya, V.; Varghese Gupta, S. Salts of Therapeutic Agents: Chemical, Physicochemical, and Biological Considerations. Molecules 2018, 23, 1719. [Google Scholar] [CrossRef]
  33. Feng, Y.; Meng, Y.; Tan, F.; Lv, L.; Li, Z.; Wang, Y.; Yang, Y.; Gong, W.; Yang, M. Effect of Surfactants and Polymers on the Dissolution Behavior of Supersaturable Tecovirimat-4-Hydroxybenzoic Acid Cocrystals. Pharmaceutics 2021, 13, 1772. [Google Scholar] [CrossRef] [PubMed]
  34. Buddhadev, S.S.; Garala, K.C. Pharmaceutical Cocrystals—A Review. Proceedings 2021, 62, 14. [Google Scholar] [CrossRef]
  35. Liu, L.; Wang, J.-R.; Mei, X. Enhancing the Stability of Active Pharmaceutical Ingredients by the Cocrystal Strategy. CrystEngComm 2022, 24, 2002–2022. [Google Scholar] [CrossRef]
  36. Yousef, M.A.E.; Vangala, V.R. Pharmaceutical Cocrystals: Molecules, Crystals, Formulations, Medicines. Cryst. Growth Des. 2019, 19, 7420–7438. [Google Scholar] [CrossRef]
  37. Duggirala, N.K.; Perry, M.L.; Almarsson, Ö.; Zaworotko, M.J. Pharmaceutical Cocrystals: Along the Path to Improved Medicines. Chem. Commun. 2015, 52, 640–655. [Google Scholar] [CrossRef]
  38. Karimi-Jafari, M.; Padrela, L.; Walker, G.M.; Croker, D.M. Creating Cocrystals: A Review of Pharmaceutical Cocrystal Preparation Routes and Applications. Cryst. Growth Des. 2018, 18, 6370–6387. [Google Scholar] [CrossRef]
  39. Kumar, A.; Nanda, A. In-Silico Methods of Cocrystal Screening: A Review on Tools for Rational Design of Pharmaceutical Cocrystals. J. Drug Deliv. Sci. Technol. 2021, 63, 102527. [Google Scholar] [CrossRef]
  40. Gadade, D.D.; Pekamwar, S.S. Pharmaceutical Cocrystals: Regulatory and Strategic Aspects, Design and Development. Adv. Pharm. Bull. 2016, 6, 479–494. [Google Scholar] [CrossRef]
  41. Chivate, A.; Garkal, A.; Dhas, N.; Mehta, T. Hot-Melt Extrusion: An Emerging Technique for Solubility Enhancement of Poorly Water-Soluble Drugs. PDA J. Pharm. Sci. Technol. 2021, 75, 357–373. [Google Scholar] [CrossRef]
  42. Shoaib, A.; Mangla, B.; Javed, S.; Sultan, M.H.; Alqahtani, S.S.; Shakeel, F. Vicissitudes of Liquid Crystals for Solubility Enhancement of Poorly Soluble Drugs. J. Mol. Liq. 2021, 321, 114924. [Google Scholar] [CrossRef]
  43. Bavishi, D.D.; Borkhataria, C.H. Spring and Parachute: How Cocrystals Enhance Solubility. Prog. Cryst. Growth Charact. Mater. 2016, 62, 1–8. [Google Scholar] [CrossRef]
  44. Zhu, Y.; Fu, Y.; Zhang, A.; Wang, X.; Zhao, Z.; Zhang, Y.; Yin, T.; Gou, J.; Wang, Y.; He, H.; et al. Rod-Shaped Nintedanib Nanocrystals Improved Oral Bioavailability through Multiple Intestinal Absorption Pathways. Eur. J. Pharm. Sci. 2022, 168, 106047. [Google Scholar] [CrossRef] [PubMed]
  45. Li, Y.-W.; Zhang, H.-M.; Cui, B.-J.; Hao, C.-Y.; Zhu, H.-Y.; Guan, J.; Wang, D.; Jin, Y.; Feng, B.; Cai, J.-H.; et al. “Felodipine-Indomethacin” Co-Amorphous Supersaturating Drug Delivery Systems: “Spring-Parachute” Process, Stability, in Vivo Bioavailability, and Underlying Molecular Mechanisms. Eur. J. Pharm. Biopharm. 2021, 166, 111–125. [Google Scholar] [CrossRef] [PubMed]
  46. Fujita, M.; Goto, S.; Chatani, H.; Otsuka, Y.; Shimada, Y.; Terada, H.; Inoo, K. The Function of Oxybuprocaine: A Parachute Effect That Sustains the Supersaturated State of Anhydrous Piroxicam Crystals. RSC Adv. 2020, 10, 1572–1579. [Google Scholar] [CrossRef] [PubMed]
  47. Nugrahani, I.; Parwati, R.D. Challenges and Progress in Nonsteroidal Anti-Inflammatory Drugs Co-Crystal Development. Molecules 2021, 26, 4185. [Google Scholar] [CrossRef]
  48. Zhang, X.; Wang, A.; Yang, X.; Wang, Y.; Wang, Q.; Hu, R.; Anwaier, G.; Di, C.; Qi, R.; Huang, Y. HPMC Improves Protective Effects of Naringenin and Isonicotinamide Co-Crystals against Abdominal Aortic Aneurysm. Cardiovasc. Drugs Ther. 2021, 1–11. [Google Scholar] [CrossRef]
  49. Ilie, A.-R.; Griffin, B.T.; Vertzoni, M.; Kuentz, M.; Kolakovic, R.; Prudic-Paus, A.; Malash, A.; Bohets, H.; Herman, J.; Holm, R. Exploring Precipitation Inhibitors to Improve in Vivo Absorption of Cinnarizine from Supersaturated Lipid-Based Drug Delivery Systems. Eur. J. Pharm. Sci. 2021, 159, 105691. [Google Scholar] [CrossRef]
  50. Malamatari, M.; Ross, S.A.; Douroumis, D.; Velaga, S.P. Experimental Cocrystal Screening and Solution Based Scale-up Cocrystallization Methods. Adv. Drug Deliv. Rev. 2017, 117, 162–177. [Google Scholar] [CrossRef]
  51. Salem, A.; Nagy, S.; Pál, S.; Széchenyi, A. Reliability of the Hansen Solubility Parameters as Co-Crystal Formation Prediction Tool. Int. J. Pharm. 2019, 558, 319–327. [Google Scholar] [CrossRef]
  52. Gunawardana, C.A.; Aakeröy, C.B. Co-Crystal Synthesis: Fact, Fancy, and Great Expectations. Chem. Commun. 2018, 54, 14047–14060. [Google Scholar] [CrossRef]
  53. Nugrahani, I.; Jessica, M.A. Amino Acids as the Potential Co-Former for Co-Crystal Development: A Review. Molecules 2021, 26, 3279. [Google Scholar] [CrossRef] [PubMed]
  54. Jagia, M.; Kale, D.P.; Bansal, A.K.; Patel, S. Novel Co-Crystals and Eutectics of Febuxostat: Characterization, Mechanism of Formation, and Improved Dissolution. AAPS PharmSciTech 2021, 23, 43. [Google Scholar] [CrossRef]
  55. Sarkar, N.; Sinha, A.S.; Aakeröy, C.B. Systematic Investigation of Hydrogen-Bond Propensities for Informing Co-Crystal Design and Assembly. CrystEngComm 2019, 21, 6048–6055. [Google Scholar] [CrossRef]
  56. Sarathi, P.; Padhi, S. Insight of the Various in Silico Screening Techniques Developed for Assortment of Cocrystal Formers and Their Thermodynamic Characterization. Drug Dev. Ind. Pharm. 2022, 47, 1523–1534. [Google Scholar] [CrossRef]
  57. MacEachern, L.; Kermanshahi-pour, A.; Mirmehrabi, M. Supercritical Carbon Dioxide for Pharmaceutical Co-Crystal Production. Cryst. Growth Des. 2020, 20, 6226–6244. [Google Scholar] [CrossRef]
  58. Cysewski, P.; Przybyłek, M. Selection of Effective Cocrystals Former for Dissolution Rate Improvement of Active Pharmaceutical Ingredients Based on Lipoaffinity Index. Eur. J. Pharm. Sci. 2017, 107, 87–96. [Google Scholar] [CrossRef] [PubMed]
  59. Etter, M.C. Hydrogen Bonds as Design Elements in Organic Chemistry. J. Phys. Chem. 1991, 95, 4601–4610. [Google Scholar] [CrossRef]
  60. Etter, M.C. Encoding and Decoding Hydrogen-Bond Patterns of Organic Compounds. Acc. Chem. Res. 1990, 23, 120–126. [Google Scholar] [CrossRef]
  61. Nikam, V.J.; Patil, S.B. Pharmaceutical Cocrystals of Nebivolol Hydrochloride with Enhanced Solubility. J. Cryst. Growth 2020, 534, 125488. [Google Scholar] [CrossRef]
  62. Yadav, S.; Gupta, C.; Sharma, N.; Kumar, J. COCRYSTALS: AN ALTERNATIVE APPROACH TO MODIFY PHYSICOCHEMICAL PROPERTIES OF DRUGS. Res. J. Pharm. Biol. Chem. Sci. 2015, 2015, 427–436. [Google Scholar]
  63. Desiraju, G.R. Supramolecular Synthons in Crystal Engineering—A New Organic Synthesis. Angew. Chem. Int. Ed. Engl. 1995, 34, 2311–2327. [Google Scholar] [CrossRef]
  64. Rajendrakumar, S.; Surampudi Venkata Sai Durga, A.; Balasubramanian, S. Strategic Synthon Approach in Obtaining Cocrystals and Cocrystal Polymorphs of a High-Z′ System Deferiprone—An Anti-Thalassemia Drug. Acta Cryst. B 2021, 77, 946–964. [Google Scholar] [CrossRef]
  65. Shi, X.; Wang, C.; Chen, Q.; Shen, S.; Song, S.; Zhou, X. Improving Physicochemical Properties of Ibrutinib with Cocrystal Strategy Based on Structures and Natures of the Carboxylic Acid Co-Formers. J. Drug Deliv. Sci. Technol. 2021, 63, 102554. [Google Scholar] [CrossRef]
  66. dos Santos, J.A.B.; Chaves Júnior, J.V.; de Araújo Batista, R.S.; de Sousa, D.P.; Ferreira, G.L.R.; de Lima Neto, S.A.; de Santana Oliveira, A.; de Souza, F.S.; Aragão, C.F.S. Preparation, Physicochemical Characterization and Solubility Evaluation of Pharmaceutical Cocrystals of Cinnamic Acid. J. Therm. Anal. Calorim. 2021, 145, 379–390. [Google Scholar] [CrossRef]
  67. Ghosh, T.; Juturu, T.; Nagar, S.N.; Kamath, S. Cocrystals of Modafinil-Nicotinic Acid: A Novel Cocrystal for Enhanced Bioavailability. Proceedings 2021, 62, 12. [Google Scholar] [CrossRef]
  68. Salem, A.; Hagymási, A.; Vörös-Horváth, B.; Šafarik, T.; Balić, T.; Szabó, P.; Gősi, F.; Nagy, S.; Pál, S.; Kunsági-Máté, S.; et al. Solvent Dependent 4-Aminosalicylic Acid-Sulfamethazine Co-Crystal Polymorph Control. Eur. J. Pharm. Sci. 2021, 156, 105599. [Google Scholar] [CrossRef]
  69. Devogelaer, J.-J.; Brugman, S.J.T.; Meekes, H.; Tinnemans, P.; Vlieg, E.; de Gelder, R. Cocrystal Design by Network-Based Link Prediction. CrystEngComm 2019, 21, 6875–6885. [Google Scholar] [CrossRef]
  70. Devogelaer, J.-J.; Meekes, H.; Tinnemans, P.; Vlieg, E.; Gelder, R. de Co-Crystal Prediction by Artificial Neural Networks**. Angew. Chem. Int. Ed. 2020, 59, 21711–21718. [Google Scholar] [CrossRef]
  71. Heng, T.; Yang, D.; Wang, R.; Zhang, L.; Lu, Y.; Du, G. Progress in Research on Artificial Intelligence Applied to Polymorphism and Cocrystal Prediction. ACS Omega 2021, 6, 15543–15550. [Google Scholar] [CrossRef]
  72. Wang, D.; Yang, Z.; Zhu, B.; Mei, X.; Luo, X. Machine-Learning-Guided Cocrystal Prediction Based on Large Data Base. Cryst. Growth Des. 2020, 20, 6610–6621. [Google Scholar] [CrossRef]
  73. Wang, L.; Yan, Y.; Zhang, X.; Zhou, X. Novel Pharmaceutical Cocrystal of Lenalidomide with Nicotinamide: Structural Design, Evaluation, and Thermal Phase Transition Study. Int. J. Pharm. 2022, 613, 121394. [Google Scholar] [CrossRef] [PubMed]
  74. Yang, S.; Liu, Q.; Ji, W.; An, Q.; Song, J.; Xing, C.; Yang, D.; Zhang, L.; Lu, Y.; Du, G. Cocrystals of Praziquantel with Phenolic Acids: Discovery, Characterization, and Evaluation. Molecules 2022, 27, 2022. [Google Scholar] [CrossRef] [PubMed]
  75. Surov, A.O.; Churakov, A.V.; Proshin, A.N.; Dai, X.-L.; Lu, T.; Perlovich, G.L. Cocrystals of a 1,2,4-Thiadiazole-Based Potent Neuroprotector with Gallic Acid: Solubility, Thermodynamic Stability Relationships and Formation Pathways. Phys. Chem. Chem. Phys. 2018, 20, 14469–14481. [Google Scholar] [CrossRef] [PubMed]
  76. Savla, P.P. The Selection and Screening of Conformers. Bombay Technol. 2021, 68, 1–8. [Google Scholar] [CrossRef]
  77. Mswahili, M.E.; Lee, M.-J.; Martin, G.L.; Kim, J.; Kim, P.; Choi, G.J.; Jeong, Y.-S. Cocrystal Prediction Using Machine Learning Models and Descriptors. Appl. Sci. 2021, 11, 1323. [Google Scholar] [CrossRef]
  78. Khan, H.W.; Elgharbawy, A.A.M.; Bustam, A.; Moniruzzaman, M. Design and Selection of Ionic Liquids Via COSMO for Pharmaceuticals and Medicine. In Application of Ionic Liquids in Drug Delivery; Goto, M., Moniruzzaman, M., Eds.; Springer: Singapore, 2021; pp. 137–164. ISBN 9789811643651. [Google Scholar]
  79. Vologzhanina, A.V. Intermolecular Interactions in Functional Crystalline Materials: From Data to Knowledge. Crystals 2019, 9, 478. [Google Scholar] [CrossRef]
  80. Wu, D.; Zhang, B.; Yao, Q.; Hou, B.; Zhou, L.; Xie, C.; Gong, J.; Hao, H.; Chen, W. Evaluation on Cocrystal Screening Methods and Synthesis of Multicomponent Crystals: A Case Study. Cryst. Growth Des. 2021, 21, 4531–4546. [Google Scholar] [CrossRef]
  81. Wang, M.; Ma, Y.; Shi, P.; Du, S.; Wu, S.; Gong, J. Similar but Not the Same: Difference in the Ability to Form Cocrystals between Nimesulide and the Pyridine Analogues. Cryst. Growth Des. 2021, 21, 287–296. [Google Scholar] [CrossRef]
  82. Barbas, R.; Kumar, V.; Vallcorba, O.; Prohens, R.; Frontera, A. Sildenafil–Resorcinol Cocrystal: XRPD Structure and DFT Calculations. Crystals 2020, 10, 1126. [Google Scholar] [CrossRef]
  83. Yang, D.; Cao, J.; Heng, T.; Xing, C.; Yang, S.; Zhang, L.; Lu, Y.; Du, G. Theoretical Calculation and Structural Analysis of the Cocrystals of Three Flavonols with Praziquantel. Cryst. Growth Des. 2021, 21, 2292–2300. [Google Scholar] [CrossRef]
  84. Khalaji, M.; Potrzebowski, M.J.; Dudek, M.K. Virtual Cocrystal Screening Methods as Tools to Understand the Formation of Pharmaceutical Cocrystals—A Case Study of Linezolid, a Wide-Range Antibacterial Drug. Cryst. Growth Des. 2021, 21, 2301–2314. [Google Scholar] [CrossRef]
  85. Sun, G.; Liu, X.; Abramov, Y.A.; Nilsson Lill, S.O.; Chang, C.; Burger, V.; Broo, A. Current State-of-the-Art In-House and Cloud-Based Applications of Virtual Polymorph Screening of Pharmaceutical Compounds: A Challenging Case of AZD1305. Cryst. Growth Des. 2021, 21, 1972–1983. [Google Scholar] [CrossRef]
  86. Dal Magro, C.; dos Santos, A.E.; Ribas, M.M.; Aguiar, G.P.S.; Volfe, C.R.B.; Lopes, M.L.L.C.; Siebel, A.M.; Müller, L.G.; Bortoluzzi, A.J.; Lanza, M.; et al. Production of Curcumin-Resveratrol Cocrystal Using Cocrystallization with Supercritical Solvent. J. Supercrit. Fluids 2021, 171, 105190. [Google Scholar] [CrossRef]
  87. Stanton, S.A.; Du, J.J.; Lai, F.; Stanton, G.; Hawkins, B.A.; Ong, J.A.; Groundwater, P.W.; Platts, J.A.; Hibbs, D.E. Understanding Hygroscopicity of Theophylline via a Novel Cocrystal Polymorph: A Charge Density Study. J. Phys. Chem. A 2021, 125, 9736–9756. [Google Scholar] [CrossRef]
  88. Haskins, M.M.; Zaworotko, M.J. Screening and Preparation of Cocrystals: A Comparative Study of Mechanochemistry vs Slurry Methods. Cryst. Growth Des. 2021, 21, 4141–4150. [Google Scholar] [CrossRef] [PubMed]
  89. Earle, R.R.; Radha, G.V. Engineering Cocrystals of Paliperidone with Enhanced Solubility and Dissolution Characteristics. Arh. Farm. 2021, 71, 393–409. [Google Scholar] [CrossRef]
  90. O’Malley, C.; McArdle, P.; Erxleben, A. Formation of Salts and Molecular Ionic Cocrystals of Fluoroquinolones and α,ω-Dicarboxylic Acids. Cryst. Growth Des. 2022, 22, 3060–3071. [Google Scholar] [CrossRef]
  91. Ferreira, P.O.; de Moura, A.; de Almeida, A.C.; dos Santos, É.C.; Kogawa, A.C.; Caires, F.J. Mechanochemical Synthesis, Thermoanalytical Study and Characterization of New Multicomponent Solid Forms of Norfloxacin with Saccharin. J. Therm. Anal. Calorim. 2022, 147, 1985–1997. [Google Scholar] [CrossRef]
  92. Chaudhari, K.R.; Savjani, J.K.; Savjani, K.T.; Shah, H. Improved Pharmaceutical Properties of Ritonavir through Co-Crystallization Approach with Liquid-Assisted Grinding Method. Drug Dev. Ind. Pharm. 2022, 47, 1633–1642. [Google Scholar] [CrossRef]
  93. Pawar, N.; Saha, A.; Nandan, N.; Parambil, J.V. Solution Cocrystallization: A Scalable Approach for Cocrystal Production. Crystals 2021, 11, 303. [Google Scholar] [CrossRef]
  94. Douroumis, D.; Ross, S.A.; Nokhodchi, A. Advanced Methodologies for Cocrystal Synthesis. Adv. Drug Deliv. Rev. 2017, 117, 178–195. [Google Scholar] [CrossRef] [PubMed]
  95. Shete, A.; Murthy, S.; Korpale, S.; Yadav, A.; Sajane, S.; Sakhare, S.; Doijad, R. Cocrystals of Itraconazole with Amino Acids: Screening, Synthesis, Solid State Characterization, in Vitro Drug Release and Antifungal Activity. J. Drug Deliv. Sci. Technol. 2015, 28, 46–55. [Google Scholar] [CrossRef]
  96. Saganowska, P.; Wesolowski, M. DSC as a Screening Tool for Rapid Co-Crystal Detection in Binary Mixtures of Benzodiazepines with Co-Formers. J. Therm. Anal. Calorim. 2018, 133, 785–795. [Google Scholar] [CrossRef]
  97. Kumar, A.; Singh, P.; Nanda, A. Hot Stage Microscopy and Its Applications in Pharmaceutical Characterization. Appl. Microsc. 2020, 50, 12. [Google Scholar] [CrossRef] [PubMed]
  98. Park, B.; Yoon, W.; Yun, J.; Ban, E.; Yun, H.; Kim, A. Emodin-Nicotinamide (1:2) Cocrystal Identified by Thermal Screening to Improve Emodin Solubility. Int. J. Pharm. 2019, 557, 26–35. [Google Scholar] [CrossRef] [PubMed]
  99. Berry, D.J.; Seaton, C.C.; Clegg, W.; Harrington, R.W.; Coles, S.J.; Horton, P.N.; Hursthouse, M.B.; Storey, R.; Jones, W.; Friščić, T.; et al. Applying Hot-Stage Microscopy to Co-Crystal Screening: A Study of Nicotinamide with Seven Active Pharmaceutical Ingredients. Cryst. Growth Des. 2008, 8, 1697–1712. [Google Scholar] [CrossRef]
  100. Zhou, Z.; Chan, H.M.; Sung, H.H.-Y.; Tong, H.H.Y.; Zheng, Y. Identification of New Cocrystal Systems with Stoichiometric Diversity of Salicylic Acid Using Thermal Methods. Pharm. Res. 2016, 33, 1030–1039. [Google Scholar] [CrossRef]
  101. Pal, S.; Roopa, B.N.; Abu, K.; Manjunath, S.G.; Nambiar, S. Thermal Studies of Furosemide–Caffeine Binary System That Forms a Cocrystal. J. Therm. Anal. Calorim. 2014, 115, 2261–2268. [Google Scholar] [CrossRef]
  102. Borkar, D.; Puranik, P.; Trivedi, H. Experimental Design Approach for Development of Cocrystals and Immediate Release Cocrystal Tablet of Atorvastatin Calcium for Enhancement of Solubility and Dissolution. J. Res. Pharm. 2020, 24, 720–737. [Google Scholar] [CrossRef]
  103. Karagianni, A.; Malamatari, M.; Kachrimanis, K. Pharmaceutical Cocrystals: New Solid Phase Modification Approaches for the Formulation of APIs. Pharmaceutics 2018, 10, 18. [Google Scholar] [CrossRef] [Green Version]
  104. Hossain Mithu, M.S.; Ross, S.A.; Hurt, A.P.; Douroumis, D. Effect of Mechanochemical Grinding Conditions on the Formation of Pharmaceutical Cocrystals and Co-Amorphous Solid Forms of Ketoconazole—Dicarboxylic Acid. J. Drug Deliv. Sci. Technol. 2021, 63, 102508. [Google Scholar] [CrossRef]
  105. Myz, S.A.; Mikhailenko, M.A.; Mikhailovskaya, A.V.; Bulina, N.V.; Gerasimov, K.B.; Politov, A.A.; Kuznetsova, S.A.; Shakhtshneider, T.P. Cocrystals of Betulin with Adipic Acid: Preparation and Thermal Behavior. J. Therm. Anal. Calorim. 2021, 147, 8235–8242. [Google Scholar] [CrossRef]
  106. Medina, C.; Daurio, D.; Nagapudi, K.; Alvarez-Nunez, F. Manufacture of Pharmaceutical Co-crystals Using Twin Screw Extrusion: A Solvent-less and Scalable Process. J. Pharm. Sci. 2010, 99, 1693–1696. [Google Scholar] [CrossRef] [PubMed]
  107. Gajda, M.; Nartowski, K.P.; Pluta, J.; Karolewicz, B. Continuous, One-Step Synthesis of Pharmaceutical Cocrystals via Hot Melt Extrusion from Neat to Matrix-Assisted Processing—State of the Art. Int. J. Pharm. 2019, 558, 426–440. [Google Scholar] [CrossRef] [PubMed]
  108. Wicaksono, Y.; Setyawan, D. Siswandono Formation of Ketoprofen-Malonic Acid Cocrystal by Solvent Evaporation Method. Indones. J. Chem. 2017, 17, 161–166. [Google Scholar] [CrossRef]
  109. Przybyłek, M.; Ziółkowska, D.; Kobierski, M.; Mroczyńska, K.; Cysewski, P. Utilization of Oriented Crystal Growth for Screening of Aromatic Carboxylic Acids Cocrystallization with Urea. J. Cryst. Growth 2016, 433, 128–138. [Google Scholar] [CrossRef]
  110. Urano, M.; Kitahara, M.; Kishi, K.; Goto, E.; Tagami, T.; Fukami, T.; Ozeki, T. Physical Characteristics of Cilostazol–Hydroxybenzoic Acid Cocrystals Prepared Using a Spray Drying Method. Crystals 2020, 10, 313. [Google Scholar] [CrossRef]
  111. Yang, J.; Hong, B.; Wang, N.; Li, X.; Huang, X.; Bao, Y.; Xie, C.; Hao, H. Thermodynamics and Molecular Mechanism of the Formation of the Cocrystals of P-Hydroxybenzoic Acid and Glutaric Acid. CrystEngComm 2019, 21, 6374–6381. [Google Scholar] [CrossRef]
  112. Vaksler, Y.A.; Benedis, D.; Dyshin, A.A.; Oparin, R.D.; Correia, N.T.; Capet, F.; Shishkina, S.V.; Kiselev, M.G.; Idrissi, A. Spectroscopic Characterization of Single Co-Crystal of Mefenamic Acid and Nicotinamide Using Supercritical CO2. J. Mol. Liq. 2021, 334, 116117. [Google Scholar] [CrossRef]
  113. Gaddam, M.; Ravouru, N. A Crystal Engineering Design to Enhance the Solubility, Dissolution, Stability and Micrometric Properties of Omeprazole via Co-Crystallization Techniques. Res. J. Pharm. Technol. 2021, 14, 356–362. [Google Scholar] [CrossRef]
  114. Kumar, R.; Thakur, A.K.; Banerjee, N.; Chaudhari, P. A Critical Review on the Particle Generation and Other Applications of Rapid Expansion of Supercritical Solution. Int. J. Pharm. 2021, 608, 121089. [Google Scholar] [CrossRef]
  115. Liu, G.; Li, J.; Deng, S. Applications of Supercritical Anti-Solvent Process in Preparation of Solid Multicomponent Systems. Pharmaceutics 2021, 13, 475. [Google Scholar] [CrossRef] [PubMed]
  116. Huang, Z. Development and Evaluation of Nano-Formulations for Immediate Release Oral Dosage Forms of Poorly Soluble Drugs. Ph.D. Thesis, Freie Universitaet Berlin, Berlin, Germany, 2022. [Google Scholar] [CrossRef]
  117. Sellers, M.S.; Lísal, M.; Brennan, J.K. Free-Energy Calculations Using Classical Molecular Simulation: Application to the Determination of the Melting Point and Chemical Potential of a Flexible RDX Model. Phys. Chem. Chem. Phys. 2016, 18, 7841–7850. [Google Scholar] [CrossRef] [PubMed]
  118. Bruni, G.; Maggi, L.; Mustarelli, P.; Sakaj, M.; Friuli, V.; Ferrara, C.; Berbenni, V.; Girella, A.; Milanese, C.; Marini, A. Enhancing the Pharmaceutical Behavior of Nateglinide by Cocrystallization: Physicochemical Assessment of Cocrystal Formation and Informed Use of Differential Scanning Calorimetry for Its Quantitative Characterization. J. Pharm. Sci. 2019, 108, 1529–1539. [Google Scholar] [CrossRef] [PubMed]
  119. Chen, J.; Wang, X.; Hong, M.; Yi, D.; Qi, M.; Ren, G. Structure Properties of Scoparone: Polymorphs and Cocrystals. J. Mol. Struct. 2019, 1191, 323–336. [Google Scholar] [CrossRef]
  120. Budziak-Wieczorek, I.; Maciołek, U. Synthesis and Characterization of a (−)-Epicatechin and Barbituric Acid Cocrystal: Single-Crystal X-Ray Diffraction and Vibrational Spectroscopic Studies. ACS Omega 2021, 6, 8199–8209. [Google Scholar] [CrossRef]
  121. Cerreia Vioglio, P.; Chierotti, M.R.; Gobetto, R. Pharmaceutical Aspects of Salt and Cocrystal Forms of APIs and Characterization Challenges. Adv. Drug Deliv. Rev. 2017, 117, 86–110. [Google Scholar] [CrossRef]
  122. Wei, H.; Gao, N.; Dang, L. Solubility and Thermodynamic Properties of Sulfamethazine–Saccharin Cocrystal in Pure and Binary (Acetonitrile + 2-Propanol) Solvents. Trans. Tianjin Univ. 2021, 27, 460–472. [Google Scholar] [CrossRef]
  123. Arabiani, M.R.; Lodagekar, A.; Yadav, B.; Chavan, R.B.; Shastri, N.R.; Purohit, P.Y.; Shelat, P.; Dave, D. Mechanochemical Synthesis of Brexpiprazole Cocrystals to Improve Its Pharmaceutical Attributes. CrystEngComm 2019, 21, 800–806. [Google Scholar] [CrossRef]
  124. Waterman, K.C.; Waterman, A.K.; Botoy, T.M.; Li, J.; Qiu, F.; Hawley, M. Stability Screening of Pharmaceutical Cocrystals. Pharm. Dev. Technol. 2021, 26, 1130–1135. [Google Scholar] [CrossRef]
  125. Alfarouk, K.O.; Stock, C.-M.; Taylor, S.; Walsh, M.; Muddathir, A.K.; Verduzco, D.; Bashir, A.H.H.; Mohammed, O.Y.; Elhassan, G.O.; Harguindey, S.; et al. Resistance to Cancer Chemotherapy: Failure in Drug Response from ADME to P-Gp. Cancer Cell Int. 2015, 15, 71. [Google Scholar] [CrossRef] [PubMed]
  126. da Silva, C.C.P.; de O. Pepino, R.; de Melo, C.C.; Tenorio, J.C.; Ellena, J. Controlled Synthesis of New 5-Fluorocytosine Cocrystals Based on the pKa Rule. Cryst. Growth Des. 2014, 14, 4383–4393. [Google Scholar] [CrossRef]
  127. He, H.; Zhang, Q.; Li, M.; Wang, J.-R.; Mei, X. Modulating the Dissolution and Mechanical Properties of Resveratrol by Cocrystallization. Cryst. Growth Des. 2017, 17, 3989–3996. [Google Scholar] [CrossRef]
  128. Duan, C.; Liu, W.; Tao, Y.; Liang, F.; Chen, Y.; Xiao, X.; Zhang, G.; Chen, Y.; Hao, C. Two Novel Palbociclib-Resorcinol and Palbociclib-Orcinol Cocrystals with Enhanced Solubility and Dissolution Rate. Pharmaceutics 2022, 14, 23. [Google Scholar] [CrossRef] [PubMed]
  129. Smith, A.J.; Kavuru, P.; Wojtas, L.; Zaworotko, M.J.; Shytle, R.D. Cocrystals of Quercetin with Improved Solubility and Oral Bioavailability. Mol. Pharm. 2011, 8, 1867–1876. [Google Scholar] [CrossRef]
  130. Iqbal, U.; Choudhary, M.I.; Yousuf, S. Synthesis of Co-Crystals of Anti-Cancer Nandrolone as a Potential Leads towards Treatment of Cancer. J. Mol. Struct. 2021, 1224, 128981. [Google Scholar] [CrossRef]
  131. Yu, Y.-M.; Liu, L.; Bu, F.-Z.; Li, Y.-T.; Yan, C.-W.; Wu, Z.-Y. A Novice Cocrystal Nanomicelle Formulation of 5-Fluorouracil with Proline: The Design, Self-Assembly and in Vitro/Vivo Biopharmaceutical Characteristics. Int. J. Pharm. 2022, 617, 121635. [Google Scholar] [CrossRef]
  132. Zhu, B.; Wang, J.-R.; Zhang, Q.; Li, M.; Guo, C.; Ren, G.; Mei, X. Stable Cocrystals and Salts of the Antineoplastic Drug Apatinib with Improved Solubility in Aqueous Solution. Cryst. Growth Des. 2018, 18, 4701–4714. [Google Scholar] [CrossRef]
  133. Yu, Y.-M.; Yu, M.-C.; Wang, L.-Y.; Li, Y.-T.; Wu, Z.-Y.; Yan, C.-W. A Supramolecular Adduct of Tegafur and Syringic Acid: The First Tegafur-Nutraceutical Cocrystal with Perfected in Vitro and in Vivo Characteristics as Well as Synergized Anticancer Activities. New J. Chem. 2020, 44, 15994–16005. [Google Scholar] [CrossRef]
  134. Li, M.; Li, Z.; Zhang, Q.; Peng, B.; Zhu, B.; Wang, J.; Liu, L.; Mei, X. Fine-Tuning the Colors of Natural Pigment Emodin with Superior Stability through Cocrystal Engineering. Cryst. Growth Des. 2018, 18, 6123–6132. [Google Scholar] [CrossRef]
  135. Shiraki, K.; Takata, N.; Takano, R.; Hayashi, Y.; Terada, K. Dissolution Improvement and the Mechanism of the Improvement from Cocrystallization of Poorly Water-Soluble Compounds. Pharm. Res. 2008, 25, 2581–2592. [Google Scholar] [CrossRef] [PubMed]
  136. Li, P.; Ramaiah, T.; Zhang, M.; Zhang, Y.; Huang, Y.; Lou, B. Two Cocrystals of Berberine Chloride with Myricetin and Dihydromyricetin: Crystal Structures, Characterization, and Antitumor Activities. Cryst. Growth Des. 2020, 20, 157–166. [Google Scholar] [CrossRef]
  137. Bolla, G.; Sarma, B.; Nangia, A.K. Crystal Engineering of Pharmaceutical Cocrystals in the Discovery and Development of Improved Drugs. Chem. Rev. 2022, 122, 11514–11603. [Google Scholar] [CrossRef] [PubMed]
  138. Xuan, B.; Wong, S.N.; Zhang, Y.; Weng, J.; Tong, H.H.Y.; Wang, C.; Sun, C.C.; Chow, S.F. Extended Release of Highly Water Soluble Isoniazid Attained through Cocrystallization with Curcumin. Cryst. Growth Des. 2020, 20, 1951–1960. [Google Scholar] [CrossRef]
  139. Wang, X.; Du, S.; Zhang, R.; Jia, X.; Yang, T.; Zhang, X. Drug–Drug Cocrystals: Opportunities and Challenges. Asian J. Pharm. Sci. 2021, 16, 307–317. [Google Scholar] [CrossRef]
  140. Yin, H.-M.; Xie, J.-Y.; Jiang, J.-Y.; Hong, M.; Zhu, B.; Ren, G.-B.; Qi, M.-H. Strategy to Tune the Performance of Two Drug Components: Drug–Drug Cocrystals of Lobaplatin with Flavonoids. Cryst. Growth Des. 2022, 22, 2602–2610. [Google Scholar] [CrossRef]
  141. da Silva, C.C.P.; de Melo, C.C.; Souza, M.S.; Diniz, L.F.; Carneiro, R.L.; Ellena, J. 5-Fluorocytosine/5-Fluorouracil Drug–Drug Cocrystal: A New Development Route Based on Mechanochemical Synthesis. J. Pharm. Innov. 2019, 14, 50–56. [Google Scholar] [CrossRef]
  142. Thorat, S.H.; Sahu, S.K.; Patwadkar, M.V.; Badiger, M.V.; Gonnade, R.G. Drug–Drug Molecular Salt Hydrate of an Anticancer Drug Gefitinib and a Loop Diuretic Drug Furosemide: An Alternative for Multidrug Treatment. J. Pharm. Sci. 2015, 104, 4207–4216. [Google Scholar] [CrossRef]
  143. Liu, L.; Li, Y.; Zhang, M.; Zhang, Y.; Lou, B. A Drug–Drug Cocrystal of Dihydromyricetin and Pentoxifylline. J. Pharm. Sci. 2022, 111, 82–87. [Google Scholar] [CrossRef]
  144. Yin, H.-M.; Wu, N.; Zhou, B.-J.; Hong, M.-H.; Zhu, B.; Qi, M.-H.; Ren, G.-B. Slow-Release Drug–Drug Cocrystals of Oxaliplatin with Flavonoids: Delaying Hydrolysis and Reducing Toxicity. Cryst. Growth Des. 2021, 21, 75–85. [Google Scholar] [CrossRef]
  145. Wang, J.; Dai, X.-L.; Lu, T.-B.; Chen, J.-M. Temozolomide–Hesperetin Drug–Drug Cocrystal with Optimized Performance in Stability, Dissolution, and Tabletability. Cryst. Growth Des. 2021, 21, 838–846. [Google Scholar] [CrossRef]
  146. Zhang, Y.; Yang, R.; Yin, H.-M.; Zhou, B.; Hong, M.; Zhu, B.; Qi, M.-H.; Ren, G.-B. Cocrystals of Flavonoids with 4,4′-Ethylenebispyridine: Crystal Structures Analysis, Dissolution Behavior, and Anti-Tumor Activity. J. Mol. Struct. 2022, 1252, 132150. [Google Scholar] [CrossRef]
  147. Yu, Y.-M.; Niu, Y.-Y.; Wang, L.-Y.; Li, Y.-T.; Wu, Z.-Y.; Yan, C.-W. Supramolecular Self-Assembly and Perfected in Vitro/Vivo Property of 5-Fluorouracil and Ferulic Acid on the Strength of Double Optimized Strategy: The First 5-Fluorouracial-Phenolic Acid Nutraceutical Cocrystal with Synergistic Antitumor Efficacy. Analyst 2021, 146, 2506–2519. [Google Scholar] [CrossRef] [PubMed]
  148. Heinen, T.; Hoelscher, S.; Vasylyeva, V. Structural Study of Anhydrous and Hydrated 5-Fluorouracil Co-Crystals with Nicotinamide and Isonicotinamide. Z. Krist. Cryst. Mater. 2022, 237, 109–116. [Google Scholar] [CrossRef]
  149. Jubeen, F.; Liaqat, A.; Amjad, F.; Sultan, M.; Iqbal, S.Z.; Sajid, I.; Khan Niazi, M.B.; Sher, F. Synthesis of 5-Fluorouracil Cocrystals with Novel Organic Acids as Coformers and Anticancer Evaluation against HCT-116 Colorectal Cell Lines. Cryst. Growth Des. 2020, 20, 2406–2414. [Google Scholar] [CrossRef]
  150. Cuadra, I.A.; Cabañas, A.; Cheda, J.A.R.; Türk, M.; Pando, C. Cocrystallization of the Anticancer Drug 5-Fluorouracil and Coformers Urea, Thiourea or Pyrazinamide Using Supercritical CO2 as an Antisolvent (SAS) and as a Solvent (CSS). J. Supercrit. Fluids 2020, 160, 104813. [Google Scholar] [CrossRef]
  151. Jubeen, F.; Liaqat, A.; Sultan, M.; Zafar Iqbal, S.; Sajid, I.; Sher, F. Green Synthesis and Biological Evaluation of Novel 5-Fluorouracil Derivatives as Potent Anticancer Agents. Saudi Pharm. J. 2019, 27, 1164–1173. [Google Scholar] [CrossRef]
  152. Gautam, M.K.; Besan, M.; Pandit, D.; Mandal, S.; Chadha, R. Cocrystal of 5-Fluorouracil: Characterization and Evaluation of Biopharmaceutical Parameters. AAPS PharmSciTech 2019, 20, 149. [Google Scholar] [CrossRef]
  153. Xu, L.-L.; Chen, J.-M.; Yan, Y.; Lu, T.-B. Improving the Solubility of 6-Mercaptopurine via Cocrystals and Salts. Cryst. Growth Des. 2012, 12, 6004–6011. [Google Scholar] [CrossRef]
  154. Yao, J.; Chen, J.-M.; Xu, Y.-B.; Lu, T.-B. Enhancing the Solubility of 6-Mercaptopurine by Formation of Ionic Cocrystal with Zinc Trifluoromethanesulfonate: Single-Crystal-to-Single-Crystal Transformation. Cryst. Growth Des. 2014, 14, 5019–5025. [Google Scholar] [CrossRef]
  155. Makadia, J.; Madu, S.J.; Arroo, R.; Seaton, C.C.; Li, M. Artemisinin–Acetylenedicarboxylic Acid Cocrystal: Screening, Structure Determination, and Physicochemical Property Characterisation. CrystEngComm 2022, 24, 1056–1067. [Google Scholar] [CrossRef]
  156. Ren, B.-Y.; Dai, X.-L.; Chen, J.-M.; Lu, T.-B. Two Anhydrous Forms and One Monohydrate of a Cocrystal of Axitinib and Glutaric Acid: Characterization, Property Evaluation and Phase Transition Study. CrystEngComm 2022, 24, 2138–2148. [Google Scholar] [CrossRef]
  157. Ren, B.-Y.; Dai, X.-L.; Wang, J.; Wu, C.; Lu, T.-B.; Chen, J.-M. Cocrystallization of Axitinib with Carboxylic Acids: Preparation, Crystal Structures and Dissolution Behavior. CrystEngComm 2021, 23, 5504–5515. [Google Scholar] [CrossRef]
  158. Pi, J.; Wang, S.; Li, W.; Kebebe, D.; Zhang, Y.; Zhang, B.; Qi, D.; Guo, P.; Li, N.; Liu, Z. A Nano-Cocrystal Strategy to Improve the Dissolution Rate and Oral Bioavailability of Baicalein. Asian J. Pharm. Sci. 2019, 14, 154–164. [Google Scholar] [CrossRef] [PubMed]
  159. Nicolov, M.; Ghiulai, R.M.; Voicu, M.; Mioc, M.; Duse, A.O.; Roman, R.; Ambrus, R.; Zupko, I.; Moaca, E.A.; Coricovac, D.E.; et al. Cocrystal Formation of Betulinic Acid and Ascorbic Acid: Synthesis, Physico-Chemical Assessment, Antioxidant, and Antiproliferative Activity. Front. Chem. 2019, 7, 92. [Google Scholar] [CrossRef]
  160. Awasthi, A.; Dheeraj, H.D.; Birangal, S.; Pai, A.; Pai, G.; Sathyanarayana, M.B. Fabrication of ceritinib cocrystals with improved solubility: Preparation, solid-state characterization, solubility studies, and molecular docking studies. RJC 2021, 14, 905–913. [Google Scholar] [CrossRef]
  161. Shahbaz, M.; Khan, U.A.; Chaudhary, M.I.; Yousuf, S. A New Bioactive Cocrystal of Coumarin-3-Carboxylic Acid and Thiourea: Detailed Structural Features and Biological Activity Studies. Acta Cryst. C 2022, 78, 192–200. [Google Scholar] [CrossRef]
  162. Paulazzi, A.R.; Alves, B.O.; Zilli, G.A.L.; dos Santos, A.E.; Petry, F.; Soares, K.D.; Danielli, L.J.; Pedroso, J.; Apel, M.A.; Aguiar, G.P.S.; et al. Curcumin and N-Acetylcysteine Cocrystal Produced with Supercritical Solvent: Characterization, Solubility, and Preclinical Evaluation of Antinociceptive and Anti-Inflammatory Activities. Inflammopharmacology 2022, 30, 327–341. [Google Scholar] [CrossRef]
  163. Pantwalawalkar, J.; More, H.; Bhange, D.; Patil, U.; Jadhav, N. Novel Curcumin Ascorbic Acid Cocrystal for Improved Solubility. J. Drug Deliv. Sci. Technol. 2021, 61, 102233. [Google Scholar] [CrossRef]
  164. Sathisaran, I.; Devidas Bhatia, D.; Vishvanath Dalvi, S. New Curcumin-Trimesic Acid Cocrystal and Anti-Invasion Activity of Curcumin Multicomponent Solids against 3D Tumor Models. Int. J. Pharm. 2020, 587, 119667. [Google Scholar] [CrossRef]
  165. Ribas, M.M.; Sakata, G.S.B.; Santos, A.E.; Dal Magro, C.; Aguiar, G.P.S.; Lanza, M.; Oliveira, J.V. Curcumin Cocrystals Using Supercritical Fluid Technology. J. Supercrit. Fluids 2019, 152, 104564. [Google Scholar] [CrossRef]
  166. Dheeraj, H.M.; Awasthi, A.; Pai, A.; Birangal, S.; Vasanthraju, S.G.; Pai, G.; Sathyanarayana, M.B. Designing of stable cocrystals of cytarabine and doxorubicin hcl using suitable coformers. Lat. Am. J. Pharm. 2021, 40, 39–48. [Google Scholar]
  167. Rai, S.K.; Gunnam, A.; Mannava, M.K.C.; Nangia, A.K. Improving the Dissolution Rate of the Anticancer Drug Dabrafenib. Cryst. Growth Des. 2020, 20, 1035–1046. [Google Scholar] [CrossRef]
  168. Chiodo, T.; Hafner, A.; Hintermann, T.; Salvador, B.; Szelagiewicz, M.; Blatter, F.; Siebenhaar, B.; Vossen, M. Multicomponent Crystals Comprising Dasatinib and Selected Cocrystal Formers. International Patent Application No. PCT/IB2013/054832, 17 May 2017. [Google Scholar]
  169. Fan, X.; Chen, J.; Shen, Q. Docetaxel–Nicotinamide Complex-Loaded Nanostructured Lipid Carriers for Transdermal Delivery. Int. J. Pharm. 2013, 458, 296–304. [Google Scholar] [CrossRef] [PubMed]
  170. Ban, E.; An, S.H.; Park, B.; Park, M.; Yoon, N.-E.; Jung, B.H.; Kim, A. Improved Solubility and Oral Absorption of Emodin-Nicotinamide Cocrystal Over Emodin with PVP as a Solubility Enhancer and Crystallization Inhibitor. J. Pharm. Sci. 2020, 109, 3660–3667. [Google Scholar] [CrossRef]
  171. Pandit, D.; Chadha, R.; Laha, B.; Gautam, M.K.; Karan, M.; Mandal, S.K. Novel Pharmaceutical Cocrystals of Gefitinib: A Credible Upswing in Strategic Research to Ameliorate Its Biopharmaceutical Challenges. Cryst. Growth Des. 2022, 22, 2218–2229. [Google Scholar] [CrossRef]
  172. Reggane, M.; Wiest, J.; Saedtler, M.; Harlacher, C.; Gutmann, M.; Zottnick, S.H.; Piechon, P.; Dix, I.; Müller-Buschbaum, K.; Holzgrabe, U.; et al. Bioinspired Co-Crystals of Imatinib Providing Enhanced Kinetic Solubility. Eur. J. Pharm. Biopharm. 2018, 128, 290–299. [Google Scholar] [CrossRef]
  173. Liu, C.; Liu, Z.; Chen, Y.; Chen, Z.; Chen, H.; Pui, Y.; Qian, F. Oral Bioavailability Enhancement of β-Lapachone, a Poorly Soluble Fast Crystallizer, by Cocrystal, Amorphous Solid Dispersion, and Crystalline Solid Dispersion. Eur. J. Pharm. Biopharm. 2018, 124, 73–81. [Google Scholar] [CrossRef]
  174. Song, J.-X.; Yan, Y.; Yao, J.; Chen, J.-M.; Lu, T.-B. Improving the Solubility of Lenalidomide via Cocrystals. Cryst. Growth Des. 2014, 14, 3069–3077. [Google Scholar] [CrossRef]
  175. Song, J.-X.; Chen, J.-M.; Lu, T.-B. Lenalidomide–Gallic Acid Cocrystals with Constant High Solubility. Cryst. Growth Des. 2015, 15, 4869–4875. [Google Scholar] [CrossRef]
  176. Hong, M.; Li, S.; Ji, W.; Qi, M.-H.; Ren, G. Cocrystals of Lenvatinib with Sulfamerazine and Salicylic Acid: Crystal Structure, Equilibrium Solubility, Stability Study, and Anti-Hepatoma Activity. Cryst. Growth Des. 2021, 21, 3714–3727. [Google Scholar] [CrossRef]
  177. Luo, Y.; Chen, S.; Zhou, J.; Chen, J.; Tian, L.; Gao, W.; Zhang, Y.; Ma, A.; Li, L.; Zhou, Z. Luteolin Cocrystals: Characterization, Evaluation of Solubility, Oral Bioavailability and Theoretical Calculation. J. Drug Deliv. Sci. Technol. 2019, 50, 248–254. [Google Scholar] [CrossRef]
  178. Hisada, N.; Takano, R.; Takata, N.; Shiraki, K.; Ueto, T.; Tanida, S.; Kataoka, M.; Yamashita, S. Characterizing the Dissolution Profiles of Supersaturable Salts, Cocrystals, and Solvates to Enhance in Vivo Oral Absorption. Eur. J. Pharm. Biopharm. 2016, 103, 192–199. [Google Scholar] [CrossRef]
  179. Bofill, L.; Barbas, R.; de Sande, D.; Font-Bardia, M.; Ràfols, C.; Albertí, J.; Prohens, R. A Novel, Extremely Bioavailable Cocrystal of Pterostilbene. Cryst. Growth Des. 2021, 21, 2315–2323. [Google Scholar] [CrossRef]
  180. Jia, J.-L.; Dai, X.-L.; Che, H.-J.; Li, M.-T.; Zhuang, X.-M.; Lu, T.-B.; Chen, J.-M. Cocrystals of Regorafenib with Dicarboxylic Acids: Synthesis, Characterization and Property Evaluation. CrystEngComm 2021, 23, 653–662. [Google Scholar] [CrossRef]
  181. Pessoa, A.S.; Aguiar, G.P.S.; Vladimir Oliveira, J.; Bortoluzzi, A.J.; Paulino, A.; Lanza, M. Precipitation of Resveratrol-Isoniazid and Resveratrol-Nicotinamide Cocrystals by Gas Antisolvent. J. Supercrit. Fluids 2019, 145, 93–102. [Google Scholar] [CrossRef]
  182. Babu, N.J.; Sanphui, P.; Nangia, A. Crystal Engineering of Stable Temozolomide Cocrystals. Chem. Asian J. 2012, 7, 2274–2285. [Google Scholar] [CrossRef]
  183. Varma, A.; Laxmi, P.; Pai, A.; Pai, G.; Sg, V.; Badamane Sathyanarayana, M. Designing of Stable Co-Crystals of Zoledronic Acid Using Suitable Coformers. Chem. Pharm. Bull. 2019, 67, 816–823. [Google Scholar] [CrossRef]
  184. Wang, L.; Li, S.; Xu, X.; Xu, X.; Wang, Q.; Li, D.; Zhang, H. Drug–Drug Cocrystals of Theophylline with Quercetin. J. Drug Deliv. Sci. Technol. 2022, 70, 103228. [Google Scholar] [CrossRef]
  185. Liu, C.; Wang, C.; Wan, S.; Liu, L.; Sun, C.C.; Qian, F. An Elusive Drug–Drug Cocrystal Prepared Using a Heteroseeding Strategy. Cryst. Growth Des. 2021, 21, 5659–5668. [Google Scholar] [CrossRef]
  186. Xiao, Y.; Yang, W.; Zhou, L.; Hao, H.; Bao, Y.; Yin, Q.; Xie, C. Growth Mechanism of the Spherulitic Propylthiouracil–Kaempferol Cocrystal: New Perspectives into Surface Nucleation. CrystEngComm 2021, 23, 2367–2375. [Google Scholar] [CrossRef]
  187. George, C.P.; Thorat, S.H.; Shaligram, P.S.; Suresha, P.R.; Gonnade, R.G. Drug–Drug Cocrystals of Anticancer Drugs Erlotinib–Furosemide and Gefitinib–Mefenamic Acid for Alternative Multi-Drug Treatment. CrystEngComm 2020, 22, 6137–6151. [Google Scholar] [CrossRef]
  188. Lv, W.-T.; Liu, X.-X.; Dai, X.-L.; Long, X.-T.; Chen, J.-M. A 5-Fluorouracil–Kaempferol Drug–Drug Cocrystal: A Ternary Phase Diagram, Characterization and Property Evaluation. CrystEngComm 2020, 22, 8127–8135. [Google Scholar] [CrossRef]
  189. Laxmi, P.; Varma, A.; Pai, A.; Sathyanarayana, M.B. Experimental Data of Fabricated Co-Crystals of Doxorubicin HCl with Flavonoids. Indian J. Pharm. Educ. Res. 2019, 53, s225–s230. [Google Scholar] [CrossRef]
  190. Deng, Y.; Zhang, Y.; Huang, Y.; Zhang, M.; Lou, B. Preparation, Crystal Structures, and Oral Bioavailability of Two Cocrystals of Emodin with Berberine Chloride. Cryst. Growth Des. 2018, 18, 7481–7488. [Google Scholar] [CrossRef]
  191. Veverka, M.; Šimon, P.; Gallovič, J.; Jorík, V.; Veverková, E.; Dubaj, T. Imatinib Mesylate Cocrystals: Synthesis, Screening, and Preliminary Characterization. Mon. Chem 2012, 143, 1405–1415. [Google Scholar] [CrossRef]
  192. Veverka, M.; Dubaj, T.; Gallovič, J.; Jorík, V.; Veverková, E.; Danihelová, M.; Šimon, P. Cocrystals of Quercetin: Synthesis, Characterization, and Screening of Biological Activity. Mon. Chem 2015, 146, 99–109. [Google Scholar] [CrossRef]
  193. Albrecht, W.; Geier, J.; RABE, S.; Palacios, D.P. Co-Crystals of Ibrutinib with Carboxylic Acids. U.S. Patent 20180072737A1, 13 August 2019. [Google Scholar]
  194. Ross, S.A.; Lamprou, D.A.; Douroumis, D. Engineering and Manufacturing of Pharmaceutical Co-Crystals: A Review on Solvent-Free Manufacturing Technologies. Chem. Commun. 2016, 52, 8772–8786. [Google Scholar] [CrossRef]
  195. Bolla, G.; Nangia, A. Pharmaceutical Cocrystals: Walking the Talk. Chem. Commun. 2016, 52, 8342–8360. [Google Scholar] [CrossRef]
  196. Wang, Y.; Zhu, W.; Du, W.; Liu, X.; Zhang, X.; Dong, H.; Hu, W. Cocrystals Strategy towards Materials for Near-Infrared Photothermal Conversion and Imaging. Angew. Chem. Int. Ed. 2018, 57, 3963–3967. [Google Scholar] [CrossRef]
Figure 1. Facts and Statistics of several cancers.
Figure 1. Facts and Statistics of several cancers.
Crystals 12 01337 g001
Figure 2. Publications on pharmaceutical cocrystals over the years.
Figure 2. Publications on pharmaceutical cocrystals over the years.
Crystals 12 01337 g002
Figure 3. Advantages of pharmaceutical cocrystals.
Figure 3. Advantages of pharmaceutical cocrystals.
Crystals 12 01337 g003
Figure 4. To generate higher apparent solubility for aqueous insoluble drugs, the spring and parachute idea was used.
Figure 4. To generate higher apparent solubility for aqueous insoluble drugs, the spring and parachute idea was used.
Crystals 12 01337 g004
Figure 5. Common supramolecular synthons formed from carboxylic acids, amides, pyridines, and other aromatic nitrogen.
Figure 5. Common supramolecular synthons formed from carboxylic acids, amides, pyridines, and other aromatic nitrogen.
Crystals 12 01337 g005
Figure 6. Conventional approaches for preparing pharmaceutical cocrystals.
Figure 6. Conventional approaches for preparing pharmaceutical cocrystals.
Crystals 12 01337 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kara, D.D.; Rathnanand, M. Cocrystals and Drug–Drug Cocrystals of Anticancer Drugs: A Perception towards Screening Techniques, Preparation, and Enhancement of Drug Properties. Crystals 2022, 12, 1337. https://doi.org/10.3390/cryst12101337

AMA Style

Kara DD, Rathnanand M. Cocrystals and Drug–Drug Cocrystals of Anticancer Drugs: A Perception towards Screening Techniques, Preparation, and Enhancement of Drug Properties. Crystals. 2022; 12(10):1337. https://doi.org/10.3390/cryst12101337

Chicago/Turabian Style

Kara, Divya Dhatri, and Mahalaxmi Rathnanand. 2022. "Cocrystals and Drug–Drug Cocrystals of Anticancer Drugs: A Perception towards Screening Techniques, Preparation, and Enhancement of Drug Properties" Crystals 12, no. 10: 1337. https://doi.org/10.3390/cryst12101337

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop