Next Article in Journal
Antiplasmodial Activity of Hydroalcoholic Extract from Jucá (Libidibia ferrea) Pods
Next Article in Special Issue
Layer-by-Layer Nanoassemblies for Vaccination Purposes
Previous Article in Journal
Exploring the Fate of Antibody-Encoding pDNA after Intramuscular Electroporation in Mice
Previous Article in Special Issue
Modified Linear Peptides Effectively Silence STAT-3 in Breast Cancer and Ovarian Cancer Cell Lines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cocrystals by Design: A Rational Coformer Selection Approach for Tackling the API Problems

by
Maan Singh
1,†,
Harsh Barua
2,†,
Vaskuri G. S. Sainaga Jyothi
3,
Madhukiran R. Dhondale
1,
Amritha G. Nambiar
1,
Ashish K. Agrawal
1,
Pradeep Kumar
4,
Nalini R. Shastri
5,* and
Dinesh Kumar
1,*
1
Pharmaceutical Solid State Research Laboratory, Department of Pharmaceutical Engineering and Technology, Indian Institute of Technology (Banaras Hindu University), Varanasi 221005, India
2
Solid State Pharmaceutical Cluster (SSPC), Science Foundation Ireland Research Centre for Pharmaceuticals, Bernal Institute, Department of Chemical Sciences, University of Limerick, V94T9PX Limerick, Ireland
3
Department of Pharmaceutics, National Institute of Pharmaceutical Education and Research, Hyderabad 500037, India
4
Wits Advanced Drug Delivery Platform Research Unit, Department of Pharmacy and Pharmacology, School of Therapeutic Sciences, Faculty of Health Sciences, University of the Witwatersrand, Johannesburg 2193, South Africa
5
Solid State Pharmaceutical Research, Hyderabad 500037, India
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Pharmaceutics 2023, 15(4), 1161; https://doi.org/10.3390/pharmaceutics15041161
Submission received: 2 March 2023 / Revised: 25 March 2023 / Accepted: 31 March 2023 / Published: 6 April 2023
(This article belongs to the Special Issue Innovative Drug Release and Vaccine Delivery Systems)

Abstract

:
Active pharmaceutical ingredients (API) with unfavorable physicochemical properties and stability present a significant challenge during their processing into final dosage forms. Cocrystallization of such APIs with suitable coformers is an efficient approach to mitigate the solubility and stability concerns. A considerable number of cocrystal-based products are currently being marketed and show an upward trend. However, to improve the API properties by cocrystallization, coformer selection plays a paramount role. Selection of suitable coformers not only improves the drug’s physicochemical properties but also improves the therapeutic effectiveness and reduces side effects. Numerous coformers have been used till date to prepare pharmaceutically acceptable cocrystals. The carboxylic acid-based coformers, such as fumaric acid, oxalic acid, succinic acid, and citric acid, are the most commonly used coformers in the currently marketed cocrystal-based products. Carboxylic acid-based coformers are capable of forming the hydrogen bond and contain smaller carbon chain with the APIs. This review summarizes the role of coformers in improving the physicochemical and pharmaceutical properties of APIs, and deeply explains the utility of afore-mentioned coformers in API cocrystal formation. The review concludes with a brief discussion on the patentability and regulatory issues related to pharmaceutical cocrystals.

Graphical Abstract

1. Introduction

In pharmaceutical research and development portfolios, about 40% of commercialized APIs exhibit low water solubility. As stated by the biopharmaceutical classification system (BCS), drugs having low solubility and high permeability fall under class II [1]. One of the most critical challenges of BCS II drugs is to improve the solubility and dissolution rate [2]. In addition to poor aqueous solubility, most APIs exhibit undesirable physicochemical properties, flowability, compactability, etc., which hampers the solid dosage form development. The purity and performance of a drug product is severely impacted due to instabilities caused by polymorphic changes and degradation due to heat, light, and humidity during processing/storage [3,4]. Among the said problems, the poor aqueous solubility of active pharmaceutical ingredients (APIs) can be enhanced by micronization, amorphization, salt formation, cocrystallization, etc. [1]. Micronization, amorphization, and salt formation can improve the aqueous solubility of drugs, but the stability and processability are compromised in some cases [5,6]. The cocrystals approach has the potential to provide a safe way to improve solubility in addition to increasing/retaining the stability [7,8].
Cocrystals can be considered superior to amorphous API forms or solid dispersions because they possess the solubility advantages of high-energy solids and have a crystalline structure with good thermodynamic stability [9,10]. Cocrystals are defined by the European Medicines Agency (EMA) as “homogenous (single phase) crystalline structures made up of two or more components in a definite stoichiometric ratio where the arrangement in the crystal lattice is not based on ionic bonds (as with salts) and the components of a cocrystal may nevertheless be neutral as well as ionized” [11]. The United States Food and Drug Administration (USFDA) defines cocrystals as “Crystalline materials composed of two or more different molecules, typically API and cocrystal formers (coformers), in the same crystal lattice in a defined stoichiometric ratio.” Cocrystals are different from salts, polymorphs, solvates and hydrates [12]. The hydrogen-bonding interactions of API with the coformer alter its physicochemical properties and lead to enhanced pharmaceutical attributes [13]. Various methods [14] have been utilized till the present by researchers to prepare pharmaceutically acceptable cocrystals and have been illustrated in Figure 1.
Commercialized drug products provide some evidence of the efficacious application of cocrystallization in the pharmaceutical industry. Depakote®, Entresto®, Suglat®, Steglatro®, Lexapro®, ESIX-10®, Beta-chlor®, Cafcit®, Zafatek®, and Lamivudine/zidovudine Teva®, etc., are some commercially available pharmaceutical products that contain cocrystal-based APIs [15,16].
The coformer selection plays an important role in deciding the final cocrystal attributes. The coformers have the ability to modulate the API stability and solubility when prepared as a cocrystal by inducing changes in its crystal structure [1]. There have been a few studies that have reported deterioration of API properties after cocrystallization [3]. A variety of (GRAS) coformers generally regarded as safe are used to prepare pharmaceutically acceptable cocrystals [1,17,18,19]. The nature of the coformer used (acidic/basic/neutral) is known to influence the stability of the final cocrystal [1]. There have been a few instances wherein the cocrystallization technique was applied to improve the hygroscopic stability of moisture-sensitive drugs [20]. Other common instabilities such as hydrolysis, isomerization, photodegradation, etc., can also be effectively overcome by means of cocrystal preparation [21,22]. A large number of coformers with different functionalities have been used till present to prepare pharmaceutical cocrystals. The utility of chemicals as coformers depends upon the hydrogen-bonding ability of the molecules with the API. The good hydrogen-bonding strength and molecular geometry between the coformer and the API plays a vital role in the development of cocrystals [23]. According to the Etter rule, the hydrogen bond is formed if good hydrogen-bond donors and acceptors participate in the hydrogen bonding [24].
Currently available literature on coformer selection is predominantly focused on the mere cocrystal formation using different coformers. However, there is very little emphasis on selecting a coformer specifically to improve a particular aspect of an API [13,25,26,27]. In this review article, the authors discuss the coformer selection, their properties and impacts on enhancing a particular physicochemical property of an API. Authors have provided statistical analyses on most commonly used coformers. The properties, applications and recent reports on the usage of commonly used aliphatic carboxylic acid-based coformers, such as succinic acid, fumaric acid, oxalic acid, and citric acid, is discussed in detail. The marketed formulations based on these four coformers are discussed and their applicability in improving the API properties of all four coformers is compared. The patentability and regulatory factors governing the development of cocrystals is briefly discussed towards the end of the article.

2. Selection of Coformer

As mentioned earlier, coformers have a major role in the cocrystal development. The factors such as the type of functional group, pKa, their physical form, and their molecular size are to be considered during cocrystal formation using a particular coformer [28]. The coformer selection is primarily done by the experimental method and knowledge-based method. The experimental method is based on trial and error. Herein, an API is cocrystallized with empirically selected coformers and the formation of cocrystals is later confirmed by employing analytical techniques such as powder X-ray diffraction (PXRD), differential scanning calorimetry (DSC), etc. This method of cocrystal screening is thus very tedious and requires a huge amount of resources. Alternatively, various knowledge-based approaches can be put to use. Suitable coformers are being selected based on the hydrogen-bonding, pKa-based models, supramolecular synthon compatibility using the Cambridge Structure Database (CSD), lattice energy calculation, Hansen solubility parameter, thermal analysis, saturation temperature measurements, virtual cocrystal screening (using molecular electrostatic potential surfaces-MEPS), etc. [29,30]. The Hansen solubility parameter (HSP) compares the aqueous solubility of the coformer and the API, and the compounds with similar HSP have a higher probability of forming cocrystals [31]. The knowledge-based methods thus predict the formation of cocrystals even before experimentation, based on the structural features of API and coformer. Another method for selecting coformers is based on the supramolecular synthons. Supramolecular synthons are the structural units within the supermolecules that can be generated due to intermolecular interactions. Supramolecular synthons are of two types: supramolecular homosynthons with identical self-complementary functionalities and supramolecular heterosynthons with different but complementary functionalities. The heterosynthons are typically more durable. In general, amide homodimers and carboxylic acid heterosynthons are preferred [32,33]. Figure 2 shows some of the common supramolecular synthons occurring in the cocrystals.
The most preferred approaches for the selection of coformer and the generation of cocrystals are Cambridge Structural Database (CSD)-based screening, hydrogen-bond rules, and pKa rules, which are briefly discussed below.

2.1. Cambridge Structural Database (CSD)

It is possible to carry out supramolecular retrosynthetic analysis, which entails locating intermolecular units for the desired cocrystal structure, using the CSD [34]. The CSD contains crystallographic information regarding the hydrogen bonds formed between the drug and the coformer. Currently, the CSD repository contains over 1.2 million crystal structures [35,36]. Every entry in the CSD contains information on chemical structure and crystallographic data (such as space groups, lattice, symmetry, and crystal systems), crystal packing, molecular dimensions, molecular geometry, stereochemistry, structure representation, and conformational analysis [37]. Based on the understanding of geometries and preferred orientations of current intermolecular interactions, coformers can be chosen for cocrystallization with the APIs [34,38].

2.2. Hydrogen-Bond Rules

The hydrogen-bond rule is another approach for the selection of coformer. The hydrogen bond (X-H) is an attractive interaction between a hydrogen atom and an electronegative atom (X). Hydrogen bonding can occur within a molecule or between two different molecules [39]. The hydrogen bond rule provides valuable information about the favored hydrogen-bond selectivity, connectivity patterns, and stereo-electronic properties of hydrogen bonds for a specific functional group or combination of functional groups, in which hydrogen bonds are formed. There are generally three rules of hydrogen bond formation. The first rule was proposed by Donohue: all available acidic hydrogen in the molecular crystal structure of that compound will be used in hydrogen bonding. The second rule states that if hydrogen-bond donors are present, all good acceptors will be engaged in hydrogen bonding. According to the third rule, hydrogen bonds will form especially between the finest hydrogen-bond acceptor and the finest hydrogen-bond donor [40].

2.3. pKa Rule

The pKa difference between the acid-base pair can be estimated to predict whether the pair forms salt or cocrystal [41]. The acid dissociation constant (pKa) affects how well certain medications are absorbed orally. The BCS Class II drugs are further classified into Iia (acidic drugs), Iib (basic drugs), and Iic (neutral drugs) based on their pH-dependent solubility (and dissolution rate). Weakly acidic drugs with pKa ≤ 5 (flurbiprofen, ketoprofen, etc.) exhibit higher aqueous solubility at alkaline pH of the intestine and are classified as class Iia drugs. The weakly basic drugs with pKa ≥ 6 (carbamazepine, rifampicin, etc.), in contrast, exhibit higher aqueous solubility at acidic pH of the stomach and are classified as class Iib drugs. Drugs which do not exhibit pH-dependent solubility are classified into the class Iic category (neutral drugs such as danazol, fenofibrate, etc.) [18,42,43]. The formation of cocrystals and salt can be predicted by a study of the transfer of protons and is determined by the ∆pKa = [pKa (base) − pKa (acid)]. A pKa value difference of greater than 2 or 3 between the API and coformer symbolizes the transfer of proton between acid and base. The smaller pKa value difference (less than 0) indicates cocrystal formation whereas a large difference in the pKa values (≥ 2 or 3) indicates salt formation [13].

3. Coformer Impact on Pharmaceutical Attributes

Coformers, along with drugs, have an ability to alter the pharmaceutical attributes of a cocrystal system. Thus, when designing a cocrystal, it is crucial to consider the physicochemical properties of coformers as well. The applicability of coformers to improve the stability, mechanical properties, solubility, and permeability of APIs is discussed in the current section. Figure 3 and Figure 4 depict the different aspects of API and coformers that need to be considered during cocrystal preparation.

3.1. Role of Coformers in Solubility, Dissolution, and Bioavailability

Oral delivery of drugs is the most favored and patient-compliant method of drug administration in spite of other available routes such as parenteral, pulmonary, transdermal, etc. The oral route offers a painless method of drug administration with high acceptance that makes it the most convenient route. Most of the APIs can be formulated as an oral solid dosage at comparatively low cost, which in totality makes them an attractive avenue for patients and pharmaceutical companies alike. The cocrystal formation effectively improves the solubility of poorly water-soluble APIs, and the coformer selected has a major role to play in deciding the solubility of the cocrystal prepared. The physicochemical properties of coformers, such as solubility, ionization, etc., are to be considered while designing cocrystals of poorly water-soluble APIs. A few examples related to the mentioned coformer properties are discussed below.
Putra et al. [21] reported cocrystals of epalrestat with betaine in which they depicted a two-fold increase in solubility. The reason was attributed to the formation of a layered structure between drug and coformer, as depicted in Figure 5. As compared to epalrestat, betaine has higher water solubility with a higher tendency to go into solution, resulting in the formation of cocrystal with higher solubility. On contact with water, the rate of epalrestat dissolution is accelerated due to rapid dissolution of betaine, which was supported by the cocrystal having a 3.5-fold higher intrinsic dissolution rate compared to the parent drug (shown in Figure 6).
Selection of a coformer of intermediate solubility directly influences cocrystal solubility, as it will lead to a prolonged parachute effect. This hypothesis is effectively highlighted by furosemide and 2-picolinamide sesquihydrate cocrystal, as reported by Banik et al. [44]. Sustained super-saturation levels of dissolved drug were maintained over the period of 24 h. This was attributed to the fact that the coformer 2-picolinamide has intermediate solubility, which led to gradual leaching of the drug. Such a phenomenon was described by a new term, the ‘synthon-extended-spring-parachute effect’, and is depicted in Figure 7. Further, the cocrystal with higher polarity will have higher solubility and consequently higher dissolution compared to neutral ones. Surov et al. [45] reported non-ionic cocrystal of diclofenac with theophylline where only a 1.6-fold increase in solubility was observed, whereas Nugrahani et al. [46] reported zwitterionic cocrystal of the same drug with l-proline showing a 7.69-fold increase in solubility.
The solubility and dissolution improvement of API are correlated with coformer solubility. In some cases, this correlation is not followed, which could be due to the interplay of other factors. One of the most plausible reasons attributed to such behavior is the existence of a stronger interaction between drug and coformer, leading to the formation of differently behaving cocrystal. There have been numerous examples of such cases in the literature. One such case was reported by Aitipamula et al. [47], who found that although nicotinamide had higher solubility than theophylline, cocrystal of flufenamic acid with theophylline showed higher solubility and dissolution compared to one with nicotinamide due to the existence of a stronger intermolecular interaction between the drug and nicotinamide. The effect of cocrystal selection on solubility and dissolution rate is summarized in Table 1.

3.2. Role of Coformers in Improving the Mechanical Properties of Drug Molecule

The APIs must have optimal mechanical properties to ensure easy processing of unit operations like mixing, granulation, tableting, etc. The properties relating to plasticity, elastic recovery, tensile strength, presence of slip planes, attachment energy, etc., are significantly influenced by the crystal packing of molecules in API [3,57,58]. Introduction of a coformer in the API crystal structure can either enhance, deteriorate or retain the mechanical properties [3,4]. Therefore, the mechanical properties can be tuned by the proper selection of coformer. There are several reports where the incorporation of coformer leads to improved mechanical properties of APIs [59]. Sun et al. [59] improved the tabletability of caffeine by co-crystallizing with methyl gallate. The synthesis of cocrystal leads to improvements in the compaction properties due to the presence of slip planes in the structure of cocrystal. The planes are held by the weak van der Waals forces and can be easily overcome by applying pressure. The planes slide over each other, which improves the tabletability of the cocrystal form. Similarly, Karki et al. [58] and Ahmed et al. [60] improved the tabletability of paracetamol by preparing its cocrystals. Further, the bending flexibility of probenecid was transformed from a single-component system to the multiple-component system by choosing a coformer having symmetrical hydrogen bond donor/acceptor groups [61]. The bending flexibility of probenecid was also retained by introducing a molecular spacer coformer, as reported by Nath et al. [62]. This indicates that, based on design strategy, the mechanical properties of drug molecules can be altered. Other examples of coformers that influence the mechanical properties of drugs are provided in Table 2.

3.3. Role of Coformers in Stabilizing the Drug Molecule

The stability of a drug during processing and storage is the most important concern of the pharmaceutical industry. Any changes in the polymorphic form of an API leads to changes in its physicochemical properties and can deteriorate the quality and safety of the final product. APIs face physical stability issues, such as pseudopolymorphism, polymorphic transition, hygroscopicity, and chemical stability issues such as hydrolysis, photolysis, thermal degradation, chemical transformation, dimerization, etc. These instabilities are mainly related to the crystal structure of API molecules. Cocrystals are ideal to avoid these instabilities [70]. The depiction of utility of cocrystals in improving the stability of APIs is shown in Figure 8. A few research works aimed at improving the stability of API are reported below.
Andrew et al. [71] employed cocrystallization to inhibit the hydration of anhydrous caffeine. Cocrystals of caffeine were synthesized with acid coformers, such as oxalic acid, glutaric acid, maleic acid, and malonic acid. It was reported that the stability of the cocrystal towards hydration followed the order of strength of acid groups in coformers, where the oxalic acid cocrystal was stable while the glutaric acid, which has the weakest acid groups, showed the least stability. It was hypothesized that the driving force for stabilization of cocrystals was achieved by employing the hydrogen bond donor such as the carboxylic acid group for the basic imidazole nitrogen. This effectively prevents water incorporation into the lattice of a cocrystal. Similar studies were conducted with etoricoxib as well, where stability enhancement was seen due to the strong hydrogen bonding between the coformer and API [72]. Gao et al. [22] improved the chemical stability of adefovir dipivoxil by cocrystallizing with acidic and basic coformers, saccharin and nicotinamide, respectively. The results showed that the acidic coformer enhanced the stability while the basic coformer did not stabilize it to such an extent. The study authors hypothesized three reasons for this: (1) the acidic coformer provided a micro-acidic environment to the structure, inhibiting the hydrolysis, whereas the basic coformer enhanced the degradation; (2) the acidic coformer introduced into the crystal lattice inhibits the dimerization, which later may lead to degradation; (3) the strong hydrogen bond of an acidic coformer prevents the moisture attack on the functional groups of drugs which are prone to hydrolysis.
Instabilities due to the isomerization of API can also make them unstable. This mechanism of instability was controlled by a coformer in the cocrystal of epalrestat [21]. The reaction cavity in the cocrystal was decreased compared to the drug, leading to hindered molecular motion, and thus inhibited the isomerization. This leads to improved photostability of the cocrystal compared to the drug molecule. Similarly, stability improvement of various APIs was achieved by cocrystal formation with a coformer, as summarized in Table 3.

3.4. Role of Coformers in Enhancing the Permeability of Cocrystals

Oral administration of compounds is the preferred route for the administration of medicines owing to its well-established benefits. The majority of cocrystal synthesis is done with an objective to improve oral bioavailability of drugs posing problems of either low solubility or low permeability or both. Abundant literature is available to demonstrate solubility improvement by the cocrystallization technique. On the other hand, the problem of poor permeability still remains underexplored. Permeability depends on complex factors, such as transporters, gastric transit time, intestinal epithelial metabolism and many other in vivo factors. As summarized in Table 4, the available literature indicates most of the experiments for permeability evaluation have been conducted through Franz-diffusion cells employing an artificial membrane.
There are multiple reasons for permeability enhancement in cocrystals. An increase in the concentration gradient due to high supersaturation levels of drug in solution leads to the increased flux rate (rate of drug permeation) [54]. The heterosynthon interactions in cocrystals can also improve the permeability by increasing thermodynamic activity. This feature can be attributed to the reduced packing efficiency of heterosynthons compared to the homosynthons, and as result, the packing efficiency and density of crystals reduces. Another reason could be the use of relatively lipophilic coformers that enhance the partition coefficient of drugs after cocrystallization and improve the rate of permeation across the cell membrane [78]. In rare cases, the coformer used may inhibit various drug transporter proteins, such as P-glycoprotein (P-gp). P-gp are efflux proteins present in the apical region of the small intestine and involved in the out-transportation of absorbed drug molecules. Drugs such as caffeine, stearic acid, etc., can inhibit the P-gp transporters and increase the permeation of drugs [69,79].
Dai et al. [78] employed Franz-type diffusion cells to study the permeability of 5-fluorouracil, its cocrystals and corresponding physical mixtures through a silicon membrane. The studied cocrystals were 5-fluorouracil (5-FU) with 3-hydroxybenzoic acid (1), 4-aminobenzoic acid (2), and cinnamic acid (3). Cumulative amounts of cocrystal permeated per unit area (Qn) and steady penetration rate (Js) were higher than those of the drug (Figure 9A). 5-FU physical mixtures with 3-hydroxybenzoic acid (1PM) and cinnamic acid (3PM) showed similar Qn while the physical mixture with 4-aminobenzoic acid (2PM) showed lesser cumulative diffusion compared to 5-FU (Figure 9B). The permeability of the 5-FU cocrystals was improved due to the replacement of the homosynthons formed between drug–drug by new drug–coformer heterosynthons. The weaker interaction between drug and coformer leads to rapid dissociation, which results in higher concentration of 5-FU cocrystals. Due to the higher concentration of 5-FU cocrystals, the flux rate increases, which leads to improved permeability.
Amaral et al. [79] reported cocrystal of dapsone with caffeine in which the permeability of drug, cocrystal and the corresponding physical mixture was evaluated in calu-3 human bronchial epithelial cells. The greater permeability in apical to basolateral direction was attributed to the complete dissociation of cocrystal and the absence of any crystal lattice interaction in the solution.
The discussed works do not consider factors such as transporter, gastric blood flow, metabolism inhibition, and beyond. Hence, there is a need to explore the role of coformer in improving permeability under the influence of the above-mentioned factors. In Table 4, instances of coformers enhancing the permeation of drugs are given. For better evaluation of permeability of cocrystals, there can be experimentation using the in vitro models that better correlate with an in vivo intestinal environment. In vitro methods consisting of artificial lipid membranes, such as parallel artificial membrane permeability assays (PAMPA) instead of cell-based such as Mardin–Darby canine kidney cells (MDCK), caco-2 cells, or tissue-based assays like intestinal membrane vesicles, would serve as better indicators of permeability improvement [80]. Figure 10 depicts the mechanism of permeability enhancement by cocrystals.
Table 4. Examples of studies reported on the effects of coformers on the permeability of drug molecules in cocrystals.
Table 4. Examples of studies reported on the effects of coformers on the permeability of drug molecules in cocrystals.
DrugCoformer(s)ModelEffect on PermeabilityMechanismReference
Dapsone (DAP)CaffeineCalu-3 human bronchial epithelial cellsGreater DAP permeability from apical to basolateral direction compared to DAP alone
Reduction in efflux rate, i.e., basolateral to apical direction of DAP compared to pure DAP
Possible effect of membrane transporters by caffeine and its metabolite theophylline[79]
Ethenzamide2,4-dihydroxybenzoic acidDiffusion apparatus with cellulose nitrate membraneImproved permeability of cocrystals compared to pure ethenzamideHigher solubility of cocrystal[81]
Entacapone (ETP)
  • Acetamide
  • Nicotinamide
  • Isonicotinamide
  • Pyrazinamide
  • Isoniazid
  • Theophylline
Diffusion apparatus with dialysis membrane-135Cocrystals of ETP with theophylline and pyrazinamide exhibited higher diffusion rate compared to pure drug
ETP-THP cocrystal hydrate had better diffusion rate compared to ETP-PYZ cocrystal
Higher solubility and higher permeability of the coformer (Theophylline BCS Class I and pyrazinamide BCS Class III)[82]
1,2,4-Thiadiazole derivativeVanillic acidFranz-type diffusion apparatus with regenerated cellulose membraneIncreased flux with no change in apparent permeability coefficient valuesHigher amount of dissolved drug leading to higher concentration gradient[83]
Adefovir dipivoxilStearic acidCaco-2 cells monolayersImproved cell permeabilityStearic acid acts as P-gp inhibitor[69]
Hydrochlorothiazide
  • Piperazine
  • Picolinamide
  • Tetramethyl pyrazine
  • Isoniazid
  • Malonamide
Franz diffusion apparatus with cellulose nitrate membraneHigher cumulative amount permeated with high diffusion rateHigher solubility leading to increased concentration gradient[84]
Lower cumulative amount permeated with marginal increase in initial flux
Marginal increase in diffusion rate with marginal increase in flux
Furosemide
  • Anthranilamide
  • 2,3,5,6-Tetramethylpyrazine
  • Adenine
  • Caffeine
Franz diffusion cell apparatus through a cellulose nitrate membraneHigher cumulative amount permeated with high diffusion rateDecrease in lattice energy with presence of weaker intermolecular interaction leading to increased solubility and ultimately higher concentration gradient[44]
Lower cumulative amount permeated with lower fluxHigh lattice energy contributing to poor solubility and lesser concentration in donor compartment
5-Fluorouracil
  • 4-Aminobenzoic acid
  • 3-Hydroxybenzoic acid
  • Cinnamic acid
Franz-type diffusion apparatus with a silicon membraneHigher cumulative amount permeated with high diffusion ratePresence of weaker intermolecular interaction between drug and coformer.
Lipophilicity of coformer.
[78]

4. Coformers Reported in the Literature

The discussion in the earlier sections of this review explains the significance of coformer properties in deciding the final characteristics of a cocrystal. In this section, the authors present a list of coformers reported in the literature with different APIs in Table 5. The fully exhaustive list is provided in Table S1 of the Supplementary Materials. The following table can be used as a reference for further cocrystallization experiments.

5. Coformers Used in High Demand

In the last two decades, the demand for coformers has upscaled, as depicted in Figure 11. Urea, nicotinamide, benzoic acid, etc., are in high demand compared to other coformers. The utility of carboxylic acid-based coformers such as succinic acid, citric acid, fumaric acid, and oxalic acid has also increased marginally. The carboxylic acid-based coformers are discussed in detail in the upcoming section of the review article.

6. Commercially Available Drug Products Based on Cocrystals

The ultimate aim of developing any technology/process is to make it reach the target population. Cocrystals play a vital role in improving the pharmaceutical properties of APIs. Commercialized cocrystal-based drug products are evidence of the efficacious application of cocrystallization in the pharmaceutical companies. Depakote®, Entresto®, Suglat®, Steglatro®, Lexapro®, ESIX-10®, Beta-chlor®, Cafcit®, Zafatek®, and Lamivudine/zidovudine Teva®, etc., are some commercially available pharmaceutical products that contain cocrystal-based APIs. The purpose of this section is to make the readers aware of the current scenario of the cocrystal technology.

6.1. Depakote®

Depakote® (other names Epilim, divalproex sodium, and Depakene) is used as an anti-epileptic agent that increases the level of gamma-aminobutyric acid. A depakote delayed-release tablet was approved by FDA in 1983, while ER was granted approval in 2002. Depakote contains valproic acid as an API and valproate sodium as a coformer. Valproic acid is in liquid form at room temperature, and sodium salt is highly hygroscopic. The cocrystal form of these two is less hygroscopic than the API itself [15,16,103,104,105,106].

6.2. Entresto®

Entresto® contains sacubitril and valsartan as an API in the fixed-dose combination. The Entresto is used in the treatment of symptomatic heart failure and to reduce the risk of cardiovascular death. It is available in the form of a film-coated tablet containing sacubitril and valsartan: 24/26 mg and was approved by FDA in 2015. Entresto® is a type of drug–drug cocrystal industrialized and marketed by Novartis, Basel, Switzerland. Valsartan is a neprilysin inhibitor and block angiotensin II receptor. Entresto® is the best example of a drug–drug cocrystal for the improvement of the pharmacokinetics properties of API due to cocrystallization. Valsartan shows a bioavailability enhancement of 50% in Entresto® compared to valsartan alone [15,16,107].

6.3. Suglat®

Suglat® is effective against selective SGLT2 (Sodium-Glucose Co-Transporter 2) inhibitor used in the treatment of diabetes and is available in the form of a tablet. Astellas, Tokyo, Japan and Kotobuki Pharmaceutical Co., Ltd., Nishina, Shizuoka, Japan (“Kotobuki”) discovered Suglat through research collaboration. Suglat® was launched by Astellus Pharma Inc. Tokyo, Japan and Kotobuki Pharmaceutical, Nishina, Shizuoka, Japan on 17 April 2014 in Tokyo. It is available in Suglat® tablets 25 mg and 50 mg. It is a good example of a cocrystal-based product that contains ipragliflozin as an API and L-proline as a coformer. Ipragliflozin absorbs moisture and is converted to a hydrate form under storage conditions. The cocrystallization of ipragliflozin` with L-proline imparts stability against hydrate formation [15,16,105,108].

6.4. Steglatro®

Steglatro® is indicated for treatment of insufficiently controlled type 2 diabetes mellitus in adults. It acts by inhibiting SGLT2. Steglatro® contains ertugliflozin as an API and L-pyroglutamic acid as a coformer. The daily recommended starting dose is 5 mg. If more prominent glycemic control is required, the dose of ertugliflozin can be raised in individuals who tolerate 5 mg once daily to 15 mg once daily. It was approved by US FDA in 2017 and is marketed by Pfizer, New York, United States. Cocrystallization here serves the purpose of stability enhancement of the API, because ertugliflozin exists as an unstable amorphous material. In this, ertugliflozin and L-pyroglutamic acid are used in a 1:1 ratio to enhance the stability and physicochemical properties of ertugliflozin [15,105,109,110].

6.5. Lexapro® & ESIX-10®

Lexapro® contains escitalopram oxalate, which is a selective serotonin reuptake inhibitor (SSRI) used to manage and treat major depressive and generalized anxiety disorders. Escitalopram is a pure S-enantiomer of racemic citalopram, which is also an antidepressant medication. It was approved by US FDA in 2002 and is marketed by Allergan, Dublin, Ireland. A similar example of escitalopram oxalate cocrystals is ESIX-10. It is available in the market in tablet form (10 mg). It was approved in 2009 for the treatment of anxiety and depression [15,16,111,112].

6.6. Beta-Chlor®

The existence of chloral betaine as a cocrystal was discovered only recently, in 2016, though it was first chemically synthesized. Another example is chloral betaine (Beta-chlor®), which was afterward recognized as a cocrystal in 2016. Chloral hydrate was the first sedative that was chemically synthesized in 1832. This cocrystal was made up of betaine and chloral hydrate. The cocrystals improve the thermal stability of chloral betaine compared to the pure drug substance. The melting point of chloral hydrate is 60 °C, whereas the melting point of the prepared cocrystal of chloral betaine was reported to be 120 °C [15,105].

6.7. Cafcit®

Cafcit® is another cocrystal that contains citrated caffeine or caffeine citrate. It is used to treat breathing problems in premature babies. Cafcit® shows better dissolution behavior and exhibits lower hygroscopicity than caffeine. According to X-ray diffraction studies, the cocrystal is held together by O-H···N hydrogen bonds between citric acid’s carboxylic acid groups and caffeine’s imidazole moieties. The approval for the cocrystal was obtained by HIKMA pharmaceuticals in 1999 [16,105,113].

6.8. Zafatek®

Zafatek® is a cocrystal-based tablet used as an anti-diabetes agent. It contains trelagliptin and succinic acid as API and coformer, respectively. Trelagliptin is an oral dipeptidyl peptidase IV inhibitor and was approved for use in Japan in March 2015. It is marketed by the Takeda pharmaceutical company, Tokyo, Japan [16,114].

6.9. Lamivudine/Zidovudine Teva®

Lamivudine-zidovudine cocrystal is the best example of a drug–drug cocrystal. It is indicated in antiretroviral combination therapy for the treatment of human immunodeficiency virus (HIV) infection. Lamivudine/zidovudine Teva is a generic product marketed by Teva pharmaceuticals. Zidovudine and lamivudine both have a number of hydrogen bond donor and acceptor groups. The cytosine fragment of lamivudine and the thymine fragment of zidovudine seem to be capable of forming synthons with substances that have complementary hydrogen-bonding groups [16,115,116,117].

6.10. Odomzo®

Odomzo® contains sonidegib as an active ingredient for the treatment of basal cell carcinoma. Odomzo® was approved by U.S. FDA in July 2015 and by EMA in August 2015. This is an example of cocrystals in which phosphoric acid is used as a coformer. The daily recommended dose is 200 mg of sonidegib, administered orally, and separated from the meal. Odomzo® is available in capsule form and is manufactured by Sun Pharmaceutical Industries Ltd., Mumbai, India [118,119,120].

6.11. Mayzent®

Mayzent® is used to treat multiple sclerosis. Mayzent® contains siponimod as an API and fumaric acid as a coformer in a stoichiometric ratio of 2:1. Mayzent® cocrystal is thermodynamically stable and is manufactured by Novartis, Basel, Switzerland. Mayzent® was approved by U.S. FDA in 2019 and available in tablet form [121,122,123].

6.12. Seglentis®

Seglentis® is a drug–drug cocrystal comprising celecoxib and tramadol. Seglentis® was approved by U.S. FDA in 2021 and is manufactured by Kowa pharmaceuticals, Alabama, United States. It is used in the treatment of acute pain, and the daily recommended dose of Seglentis® is 100 mg (56 mg celecoxib and 44 mg tramadol hydrochloride) [122,124].

6.13. Dimenhydrinate

Dimenhydrinate is the cocrystal of diphenhydramine (drug) and 8-chlorotheophylline (coformer). It was approved by U.S. FDA in 1982 and is manufactured by Watson Laboratories Inc, New Jersey, United States. Dimenhydrinate is available in tablet form (50 mg) for the treatment of motion sickness, including nausea and vomiting [17,125,126].

6.14. Ibrutinib

Ibrutinib, an anticancer medication used to treat chronic lymphocytic leukemia, was combined with fumaric acid to create a cocrystal that has better stability while exhibiting similar solubility to the original API. This cocrystal is still waiting for FDA clearance [15,127].

6.15. E-58425 (Clinical Trial Phase 3)

E-58425 is currently under phase 3 clinical trial for the treatment of severe acute post-operative pain. E-58425 is based on the cocrystallization technique to make the cocrystals of tramadol/celecoxib. This is an example of a drug–drug cocrystal in which two APIs are used in the formation of cocrystals. E-58425 exhibited superior analgesic activity compared to the tramadol and celecoxib combination. The synergistic effect of tramadol–celecoxib cocrystals was also reported by Manuel et al. in the treatment of severe acute post-operative pain. Sebastian et al. also reported that the co-crystal of tramadol–celecoxib (CTC) had a significant impact on efficacy in a phase 2 clinical trial [128,129,130,131,132].

6.16. TAK-020 (Clinical Trial Phase 1)

TAK-020 with Gentisic acid (coformer) is used in the treatment of rheumatoid arthritis. Currently TAK-020 is available in oral solution form. Takeda Pharmaceuticals, Tokyo, Japan is currently working on the TAK-020–gentisic acid cocrystals to form the tablet. TAK-020 is in phase 1 clinical trial. If Takeda Pharmaceuticals get positive outcomes from the clinical trials, it would be the first solid dosage form of the TAK-020. Kouya Kimoto also reported that cocrystals of TAK-020 with gentisic acid showed an enhanced dissolution rate [128,129,133,134]. All the above-mentioned cocrystal-based products are enlisted in Table 6.

7. Most Popular Coformers Utilized in Cocrystal-Based Marketed Formulations

On the basis of the available literature, aliphatic carboxylic acid-based coformers are the most commonly used in the marketed cocrystal preparations. These coformers exhibit favorable hydrogen-bonding interactions with the APIs, resulting in the formation of cocrystals. Though other coformers are reported to be used extensively in cocrystal research, the carboxylic acid-based coformers have surpassed all other coformers in terms of usage in marketed formulations. The reason for the usage of these coformers is not exactly known but they do possess a good number of hydrogen-bond donors and acceptors, which is an essential feature of cocrystal formation. Currently, the usage of coformers in marketed formulations is in the following order: fumaric acid, oxalic acid > succinic acid > citric acid. Additionally, a literature search was carried out in common search engines such as ScienceDirect, Web of Science and PubMed with the “name of coformer” followed by the word “cocrystals” as key words. The results are shown in Figure 12. From this search we can assess the current scenario of the scientific publications based on the mentioned coformers.

7.1. Fumaric Acid (FA)

The first instance of natural fumaric acid isolation was carried out from the plant fumaria officinalis. The other names of fumaric acid are trans-1,2-ethylenedicarboxylic acid or (E)-2-butenedioic acid; the term “fumarates” is also used synonymously. Chemically, fumaric acid can be synthesized from maleic anhydride. Fumaric acid is a colorless crystalline solid. The molecular formula of fumaric acid is C4H4O4. Fumaric acid is degraded by both aerobic and anaerobic microorganisms [136,137,138,139]. The chemical structure of fumaric acid is shown in Figure 13. The properties of fumaric acid along with other carboxylic acid-based coformers are tabulated in Table 7.

7.1.1. Pharmaceutical Properties

Esters of fumaric acid such as mono and dimethyl fumarate have good pharmaceutical application in the treatment of multiple sclerosis and psoriasis. In 1994, DMF was initially made available on the market as Fumaderm® [145].

7.1.2. Fumaric Acid as a Coformer

Yang et al. [7] reported that fumaric acid can be a choice of coformer to form cocrystals. Dezhi et al. reported that fumaric acid has good water solubility compared to berberine chloride (BBC). BBC possess good pharmacological activities, but poor stability limits its applications. BBC–fumaric acid cocrystals improve the stability and dissolution rate compared to BBC alone. Similarly, cocrystals of promethazine hydrochloride with the fumaric acid as coformer in the ratio 2:1 possess good solubility and stability [146]. A few research works have demonstrated the use of fumaric acid as a coformer in improving the dissolution rate (approximately 6.1 × 10−3 mmol cm−2 min−1) of fluoxetine hydrochloride [147,148]. It is also reported that fumaric acid can effectively improve the therapeutic efficacy of the APIs by improving their physicochemical properties. Enoxacin is an anti-bacterial of the fluoroquinolone class having poor aqueous solubility. The cocrystal of enoxacin with fumaric acid enhanced the solubility and permeability of the drug and thereby improved its anti-bacterial activity as well [149,150,151]. The cocrystals of 6-nitroquinoline were grown using fumaric acid as a coformer in a 1:1 ratio by using the slow solvent evaporation method. The hydrogen bonds C–H···O and O–H···N in 6-nitroquinoline fumaric acid cocrystals stabilized the structure of the API [152]. Chaitanya et al. [153] reported that the use of fumaric acid as a coformer enhances the solubility, dissolution rate, and permeability of nicorandil. They also reported that nicorandil fumaric acid cocrystals have a good hardness property at lower compaction pressure. It is also reported that single crystals of L-histidinium can be synthesized by using fumaric acid as a coformer [154]. Similarly, another report shows that fumaric acid is capable of forming single crystals with L-phenylalanine. The obtained single cocrystal of L-phenylalanine has good thermal stability [155]. The hydrogen-bonding interaction of API with fumaric acid in cocrystals is depicted in Figure 14.
According to the reported literature, fumaric acid used as a coformer in the formation of cocrystals plays a vital role. Fumaric acid showed the higher impact on the solubility and dissolution rate. Sildenafil–fumaric acid cocrystals obtained by the slow solvent evaporation method showed a great improvement in the solubility of nearly 5-fold compared to sildenafil alone. Not only did sildenafil cocrystals show improved solubility, but other cocrystals with fumaric acid also showed an increment in the solubility [156]. In the case of ketoconazole–fumaric acid cocrystals taken in the molar ratio 1:1, 1:2, and 1:3 made by using the slow solvent evaporation method, all three-molar ratio cocrystals showed improvement in the solubility, dissolution rate, and stability of the ketoconazole. The dissolution rate of ketoconazole–fumaric acid cocrystals was enhanced 1.65-fold compared to ketoconazole alone [157]. Cocrystals of promethazine hydrochloride with fumaric acid were prepared by mechanochemistry and slow solvent evaporation in the same molar ratio of 2:1. Both the promethazine hydrochloride–fumaric acid cocrystals have improved solubility and stability [146]. Below are the some reported cocrystals with fumaric acid as coformer showing improvement in solubility, dissolution rate, permeability, and stability. The utility of fumaric acid as coformer and its impact on the properties of prepared cocrystals is summarized in Table 8.

7.2. Oxalic Acid (OA)

Oxalic acid can be naturally obtained from bacteria, plants, fungi, and animals or can be chemically synthesized. Oxalic acid is odorless alpha, omega-dicarboxylic acid. The IUPAC name is ethanedioic acid and the formula is C2H2O4 [91,140,141]. The chemical structure of oxalic acid is shown in Figure 15. Refer to Table 7 for the other physical and chemical properties of the oxalic acid.

Oxalic Acid as a Coformer

From the earlier discussion, it is clear that a coformer plays a vital role in the formulation of cocrystal-based products. Oxalic acid is a coformer which has a good water solubility. Hrinova et al. [91] used oxalic acid as a coformer in the preparation of rivaroxaban cocrystals. Rivaroxaban belongs to BCS class II, having low solubility and high permeability. Oxalic acid is a highly water-soluble coformer that interacted with the rivaroxaban, forming the hydrogen bond. Rivaroxaban oxalic acid cocrystal not only increased the solubility but also showed significant improvement in the dissolution rate. Chen et al. [166] reported enhancement in solubility and bioavailability of apixaban cocrystal with oxalic acid as coformer. The prepared cocrystal performed better than the marketed product Eliquis®. Another report by Kusuma et al. [167] described improvement in stability of temozolomide which is an anti-cancer drug, marketed under the brand name Temodar® or Temodal®. Temozolomide often changes in physical appearance from white to light tan/pink during storage. This discoloration is indicative of the degradation of temozolomide. Upon cocrystallization with oxalic acid, temozolomide showed better storage stability. Similarly, escitalopram oxalate–oxalic acid cocrystals are marketed under the trade name Lexapro® in which oxalic acid is used as a coformer. The escitalopram oxalate has the stability issue, which is improved by oxalic acid after successful generation of escitalopram oxalate cocrystals [15,105]. Another report by Karki et al. [58] demonstrated that paracetamol (acetyl-para-aminophenol)–oxalic acid cocrystals are capable of improving the poor compressibility during tablet production. Chen et al. [168] reported that oxalic acid used as coformer in the preparation of xanthotoxin cocrystals enhanced its solubility, dissolution rate, and stability. The hydrogen-bonding interaction between the different APIs and oxalic acid is shown in Figure 16.
Oxalic acid is also used as a coformer to enhance the solubility, dissolution rate, permeability, stability, and bioavailability of the drugs. According to some reported literature, there was no impact on the permeability of the drug, but there was a higher impact on the solubility on the drugs. The solubility of the drug increased up to 12-fold depending on the molar ratio, method of preparation, and other factors. Xanthotoxin–oxalic acid cocrystals showed a nearly 1.6-fold solubility improvement, while rebamipide–oxalic acid cocrystals showed a 7.29-fold solubility increment. Both cocrystals were made by liquid-assisted grinding, but there was a difference in the molar ratio of drug and oxalic acid. Rebamipide–oxalic acid cocrystals also showed an improvement in bioavailability [169]. It is also reported that apixaban–oxalic acid cocrystals taken in a molar ratio of 4:3 increased solubility approximately 2-fold and bioavailability 2.7-fold [166]. Table 9 summarizes the applications of oxalic acid as coformer in improving the physicochemical properties of APIs.

7.3. Succinic Acid (SA)

Georgius Agricola, a German chemist, discovered succinic acid (also known as butanedioic) with the molecular formula C4H6O4 [173]. SA is a dicarboxylic acid that exists as white, glittering crystals [174]. The chemical structure of succinic acid is shown in Figure 17. Other physical and chemical properties of succinic acid are given in Table 7.
The global demand for succinic acid is increasing and is approximately 30,000 tons per annum [175]. Market demand for SA is increasing tremendously, from USD 131.7 million in 2018 to USD 182.8 million in 2023 at a 6.8% CAGR (compound annual growth rate) [176].
Succinic acid has a good aqueous solubility that triggers the solubility and dissolution rate of APIs that belong to BCS class II. Succinic acid is commonly used as a coformer for the preparation of cocrystals [142]. Coformers are the backbone of cocrystal formation and of their solubility enhancement. Succinic acid is soluble in water (71 mg/mL) and is used as a coformer to improve the solubility of BCS class-II drugs. By introducing a more soluble coformer into the crystal lattice, which results in a lower solvation barrier, significant attempts have been undertaken in recent decades to increase the solubility, permeability, or bioavailability of poorly water-soluble medicines. The ultimate aim of using the coformer is to improve the solubility and dissolution rate of a poorly aqueous soluble drug by introducing the coformers into the BCS class-II drugs [18,92,142].

7.3.1. Succinic Acid as a Coformer

Alhalaweh et al. [177] reported that urea-succinic acid cocrystals improve the solubility and thermodynamic stability of urea. In this work, urea as an API (e.g., to treat eczema & psoriasis) and succinic acid as a coformer was used to form the cocrystals. An acid–amide heterosynthon stabilized the 1:1 U-SA cocrystal, whereas amide–amide homosynthons and acid–amide heterosynthons stabilized the 2:1 cocrystals. The hydrogen bond interaction takes place between different APIs and succinic acid (coformer) to form the cocrystals, as shown in Figure 18.
In the case of itraconazole-succinic acid cocrystals, the dissolution rate and stability were increased compared to pure itraconazole. Two formulations were made by a liquid anti-solvent method (F1) and a gas anti-solvent method (F2). F1 achieved 50% drug release in 2 h while F2 achieved 92% drug release in 2 h. Carbamazepine–succinic acid cocrystals showed enhanced solubility, dissolution rate, stability and bioavailability compared to pure carbamazepine. Furthermore, fluoxetine–succinic acid cocrystals were made by using the slow solvent evaporation method in which fluoxetine–succinic acid was taken in the molar ratio 2:1. In this case, only the solubility of fluoxetine increased, approximately 2-fold. Table 10 below summarizes the role of succinic acid in enhancing the solubility, dissolution rate, bioavailability and stability of drugs.

7.3.2. Other Applications

Succinates (most commonly calcium succinate, potassium succinate, and sodium succinate) are extremely helpful in the treatment of long-term illnesses and injuries. These are typically employed medically as sedatives, antispasmodics, antirheumatics, and contraceptives. Succinic acid is also employed as an antioxidant and a potassium ion inhibitor. Succinic acid is also a useful product for athletes. As a result, the dicarboxylate could be considered as an “elixir of youth” [96].

7.4. Citric Acid

Citric acid is a colorless, odorless white crystalline powder; its chemical name is 2-hydroxypropane-1,2,3-tricarboxylic acid. The molecular formula of citric acid anhydrate is C6H8O7. The other physical and chemical properties of the citric acid are mentioned in the Table 7. Lemon, orange, pineapple, strawberry, red currant, cranberry, and other fruits mostly contain citric acid. In 2021, the volume of the global citric acid market was 2.7 million tons. By 2027, the market is anticipated to grow to 3.2 million tons [187,188]. The chemical structure of citric acid is shown in Figure 19.

Citric Acid as a Coformer

There are various studies showing the successful applicability of citric acid as a coformer in increasing the solubility of poorly aqueous soluble drugs by many times. Yan et al. [143] reported that the solubility of metformin HCl was increased by using citric acid as a coformer. A similar example of berberine chloride was studied by Lu et al. [6] who reported that berberine chloride shows stability issues during wet granulation for tablet production. Cocrystals of berberine chloride with citric acid as a coformer were more stable compared to berberine chloride alone. A study by Hsu et al. [189] reported that the stability of theophylline improved after preparation of its cocrystal with citric acid. Deng et al. [190] reported that dapagliflozin possess the stability problem at high temperature and also has hygroscopicity issues. Cocrystals of dapagliflozin made by the use of citric acid (coformer) improved the stability of dapagliflozin. Furthermore, Wang et al. [56] described in their research that pyrazinamide (an anti-tuberculosis drug) belongs to BCS class II and has a solubility problem. Cocrystals of pyrazinamide with citric acid enhanced both the solubility and dissolution rate compared to pyrazinamide alone. Additionally, norfloxacin-citric acid cocrystals showed improved solubility compared to norfloxacin alone [191]. Another report by Revika et al. [192] described that ethyl p-methoxycinnamate used as anti-inflammatory agent showed a 44.19% increase in solubility in its cocrystal form compared to ethyl p-methoxycinnamate alone. Fahad et al. [193] reported that cocrystals of simvastatin in which citric acid was used as a coformer showed greater solubility, dissolution rate, and bioavailability. The hydrogen-bonding interaction between the different APIs and citric acid is shown in Figure 20.
Metformin hydrochloride–citric acid cocrystals obtained by solution crystallization, neat grinding, and liquid-assisted grinding in the ratio of 1:1 increased its solubility by 1–4-fold. The bioavailability enhancement of metformin cocrystals was seen as well. Furthermore, rebamipide–citric acid cocrystals exhibited 12.58-fold improvement in solubility compared to rebamipide alone. The intrinsic dissolution rate of the cocrystal was ~13.2 times higher than the API alone [169]. It is essential to consider the molar ratio of API:coformer taken during the cocrystallization experiments [194]. For example, simvastatin–citric acid required a molar ratio of 1:1 to form cocrystals having improved the solubility profile. Conversely, nefiracetam–citric acid cocrystals required an API:coformer ratio of 2:1 to show improvement in the solubility. Various other examples have been summarized in Table 11.

8. Comparison of Coformers

On the basis of scientific papers and the literature trend, four coformers are selected: fumaric acid, oxalic acid, succinic acid, and citric acid. All the coformers used in the cocrystals showed the improvement in the solubility, dissolution rate, bioavailability, and stability, while only fumaric acid also showed the permeability enhancement. Table 12 summarizes the role of fumaric acid, oxalic acid, succinic acid, and citric acid in improving the physicochemical properties of drug.
The number of commercialized products comprised of fumaric acid or oxalic acid as a coformer are higher compared to the other two coformers. The scientific literature and approved commercialized products indicate the successful application of fumaric and oxalic acid as coformers in generating cocrystals of problematic APIs. However, there is a need for studies exploring the applicability of any particular coformer in improving the physicochemical properties of a particular class of APIs. The cocrystal formation is principally governed by the intermolecular hydrogen-bonding interactions and the changes these interactions bring about in the crystal structure. Hence, further research is required to support the claim of fumaric and oxalic acid as the best-suited coformer for API cocrystallization. Based on currently available marketed cocrystal-based formulations, it can be said that the carboxylic acid-based compounds (fumaric acid and oxalic acid in particular) can be best suited for cocrystallization to improve pharmaceutical properties. However, the authors feel that industrial input is essential to prove the potential of any coformer.

9. Patentability Issue Criteria and Regulatory Guidelines of Pharmaceutical Cocrystals

Figure 21 shows the major steps involved in the development of cocrystal-based formulations. After the successful development of a cocrystal-based product, the next stage is to obtain the necessary approvals from the concerned regulatory bodies for commercialization of the developed product. Hence, this section briefly explains the current regulatory scenario with respect to product patenting and filing.
It is necessary to improve the regulatory procedure for the filing and granting of patent as well as for the regulatory approval to commercialize the product. There are generally three conditions for the granting of patent, such as novelty, non-obviousness and utility. There are two ways to file the patent application in patent office: either by national phase application or an international phase application (PCT route). The guidelines for the pharmaceutical cocrystals were first published by the USFDA in 2013. As per the guideline, pharmaceutical cocrystals are considered to be a drug product intermediate that requires additional regulation. In the latest guidelines from USFDA in 2018, cocrystals were included as a drug substance and defined as “crystalline materials composed of two or more different molecules, one of which is the API, in a defined stoichiometric ratio within the same crystal lattice that are associated by nonionic and noncovalent bonds”. The USFDA also stated that a coformer is the component that interacts non-ionically with the API in the crystal lattice, that is not a solvent (including water), and is typically nonvolatile. For the regulatory approval of pharmaceutical cocrystals, there are two possibilities: the new drug application (NDA) pathway (505(b)(2)), and the abbreviated new drug application (ANDA) pathway (505(j)). The condition for an NDA application is that the cocrystals not have an active pharmaceutical ingredient that is already a reference listed drug (RLD) [122]. On the other hand, the applicant can file an ANDA application for the cocrystals which contains the previously approved drug (RLD). Mayzent is the best example for the newly approved cocrystals through the NDA route because its active pharmaceutical ingredient was not mentioned in the RLD [12,122,129,199].

10. Conclusions

In the current scenario, poor solubility, poor dissolution rate, poor permeability, low bioavailability, and instability are the major reasons for the failure of an active pharmaceutical ingredient (API). The researchers are focusing on mitigating these APIs issues. Cocrystallization is a proven approach to enhance the physicochemical properties of APIs and thus overcome the problems associated with APIs. Coformers play a paramount role in cocrystallization, as the final properties of cocrystal are dependent on the coformer characteristics and its interaction with the API. Research works have been presented wherein the cocrystallization experiments have resulted in improvement of the earlier-mentioned API properties. However, there is not enough evidence to prove the usage of any particular coformer to solve all API issues. A coformer must be selected based on the nature of the interaction with the API and the positive changes it brings about in the crystal structure after association. On the basis of commercialization potential, the aliphatic carboxylic acid-based coformers have gained prominence. Almost six commercialized cocrystal-based products are based on fumaric acid (two products), oxalic acid (two products), succinic acid (one product), and citric acid (one product). A few research works have been presented in this article, which shows the nature of the association of these coformers with the API and their utility in cocrystallization experiments. The ongoing research on these coformers indicates that their utility is still under exploration, which is a positive indication for cocrystallization-based research. Moreover, a small but important section related to the patentability and regulations concerning cocrystals is presented. Herein, the possible routes for filing a patent or product are discussed. Finally, the authors urge the readers/researchers to consider the carboxylic acid-based coformers for cocrystallization experiments so as to obtain concrete information on their utility as ultimate coformers.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pharmaceutics15041161/s1, Table S1: Reported coformers in literature used in the formation of cocrystals.

Author Contributions

M.S.: Conceptualization, literature survey, compilation of data, original draft writing., H.B.: Conceptualization, literature survey, compilation of data, original draft writing., V.G.S.S.J.: Writing, reviewing and editing., M.R.D.: Writing, reviewing, editing, and corrections., A.G.N.: Writing, reviewing, and editing., A.K.A.: Writing, reviewing, and editing., P.K.: Writing, reviewing, and editing., N.R.S.: Conceptualization, writing, editing, reviewing, overall modification, and correction., D.K.: Conceptualization, writing, editing, reviewing, overall modification, and correction. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful to the Indian Institute of Technology (Banaras Hindu University), Varanasi for providing infrastructure facilities. Singh, M and Kumar, D acknowledges funding from SERB, India under grant number SRG/2021/002389. Dhondale, MR acknowledges funding from PMRF, India under grant number 1102018.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article and Supplementary file.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Koranne, S.; Krzyzaniak, J.F.; Luthra, S.; Arora, K.K.; Suryanarayanan, R. Role of Coformer and Excipient Properties on the Solid-State Stability of Theophylline Cocrystals. Cryst. Growth Des. 2019, 19, 868–875. [Google Scholar] [CrossRef]
  2. Ullah, M.; Hussain, I.; Sun, C.C. The development of carbamazepine-succinic acid cocrystal tablet formulations with improved in vitro and in vivo performance. Drug Dev. Ind. Pharm. 2016, 42, 969–976. [Google Scholar] [CrossRef]
  3. Chattoraj, S.; Shi, L.; Chen, M.; Alhalaweh, A.; Velaga, S.; Sun, C.C. Origin of Deteriorated Crystal Plasticity and Compaction Properties of a 1:1 Cocrystal between Piroxicam and Saccharin. Cryst. Growth Des. 2014, 14, 3864–3874. [Google Scholar] [CrossRef]
  4. Rahman, Z.; Agarabi, C.; Zidan, A.S.; Khan, S.R.; Khan, M.A. Physico-mechanical and stability evaluation of carbamazepine cocrystal with nicotinamide. AAPS PharmSciTech 2011, 12, 693–704. [Google Scholar] [CrossRef] [Green Version]
  5. Vandana, K.R.; Prasanna Raju, Y.; Harini Chowdary, V.; Sushma, M.; Vijay Kumar, N. An overview on in situ micronization techniqueAn emerging novel concept in advanced drug delivery. Saudi Pharm. J. 2014, 22, 283–289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Lu, Q.; Dun, J.; Chen, J.M.; Liu, S.; Sun, C.C. Improving solid-state properties of berberine chloride through forming a salt cocrystal with citric acid. Int. J. Pharm. 2019, 554, 14–20. [Google Scholar] [CrossRef]
  7. Yang, D.; Cao, J.; Jiao, L.; Yang, S.; Zhang, L.; Lu, Y.; Du, G. Solubility and Stability Advantages of a New Cocrystal of Berberine Chloride with Fumaric Acid. ACS Omega 2020, 5, 8283–8292. [Google Scholar] [CrossRef] [Green Version]
  8. Rasenack, N.; Müller, B.W. Micron-size drug particles: Common and novel micronization techniques. Pharm Dev Technol 2004, 9, 1–13. [Google Scholar] [CrossRef]
  9. Issa, N.; Barnett, S.A.; Mohamed, S.; Braun, D.E.; Copley, R.C.B.; Tocher, D.A.; Price, S.L. Screening for cocrystals of succinic acid and 4-aminobenzoic acid. CrystEngComm 2012, 14, 2454–2464. [Google Scholar] [CrossRef]
  10. Zhang, T.; Yu, Q.; Li, X.; Ma, X. Preparation of 2:1 urea-succinic acid cocrystals by sublimation. J. Cryst. Growth 2017, 469, 114–118. [Google Scholar] [CrossRef]
  11. European Medicines Agency. Reflection Paper on the Use of Cocrystals of Active Substances in Medicinal Products; European Medicines Agency: Amsterdam, The Netherlands, 2015.
  12. USFDA. Regulatory Classification of Pharmaceutical Co-Crystals Guidance for Industry. 7. Available online: https://www.fda.gov/files/drugs/published/Regulatory-Classification-of-Pharmaceutical-Co-Crystals.pdf. (accessed on 22 December 2022).
  13. Childs, S.L.; Stahly, G.P.; Park, A. The salt−cocrystal continuum: The influence of crystal structure on ionization state. Mol. Pharm. 2007, 4, 323–338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. 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]
  15. Kumar Bandaru, R.; Rout, S.R.; Kenguva, G.; Gorain, B.; Alhakamy, N.A.; Kesharwani, P.; Dandela, R. Recent Advances in Pharmaceutical Cocrystals: From Bench to Market. Front Pharm. 2021, 12, 780582. [Google Scholar] [CrossRef] [PubMed]
  16. Bolla, G.; Sarma, B.; Nangia, A.K. Crystal engineering and pharmaceutical crystallization. In Hot Topics in Crystal Engineering; Elsevier: Amsterdam, The Netherlands, 2021; pp. 157–229. [Google Scholar]
  17. Cheney, M.L.; Weyna, D.R.; Shan, N.; Hanna, M.; Wojtas, L.; Zaworotko, M.J. Coformer selection in pharmaceutical cocrystal development: A case study of a meloxicam aspirin cocrystal that exhibits enhanced solubility and pharmacokinetics. J. Pharm. Sci. 2011, 100, 2172–2181. [Google Scholar] [CrossRef] [PubMed]
  18. Mondal, P.K.; Mirmehrabi, M.; Rohani, S. Insight into the Formation of Heteromolecular Hydrogen Bonds between Dasatinib and GRAS Molecules. Org. Process Res. Dev. 2021, 25, 1579–1588. [Google Scholar] [CrossRef]
  19. Wathoni, N.; Sari, W.A.; Elamin, K.M.; Mohammed, A.F.A.; Suharyani, I. A Review of Coformer Utilization in Multicomponent Crystal Formation. Molecules 2022, 27, 8693. [Google Scholar] [CrossRef]
  20. Dhondale, M.R.; Thakor, P.; Nambiar, A.G.; Singh, M.; Agrawal, A.K.; Shastri, N.R.; Kumar, D. Co-Crystallization Approach to Enhance the Stability of Moisture-Sensitive Drugs. Pharmaceutics 2023, 15, 189. [Google Scholar] [CrossRef]
  21. Putra, O.D.; Umeda, D.; Nugraha, Y.P.; Nango, K.; Yonemochi, E.; Uekusa, H. Simultaneous Improvement of Epalrestat Photostability and Solubility via Cocrystallization: A Case Study. Cryst. Growth Des. 2017, 18, 373–379. [Google Scholar] [CrossRef]
  22. Gao, Y.; Gao, J.; Liu, Z.; Kan, H.; Zu, H.; Sun, W.; Zhang, J.; Qian, S. Coformer selection based on degradation pathway of drugs: A case study of adefovir dipivoxil-saccharin and adefovir dipivoxil-nicotinamide cocrystals. Int. J. Pharm. 2012, 438, 327–335. [Google Scholar] [CrossRef]
  23. Fei, T.; Lv, P.; Liu, Y.; He, C.; Sun, C.; Pang, S. Design and Synthesis of a Series of CL-20 Cocrystals: Six-Membered Symmetrical N-Heterocyclic Compounds as Effective Coformers. Cryst. Growth Des. 2019, 19, 2779–2784. [Google Scholar] [CrossRef]
  24. Etter, M.C. Encoding and decoding hydrogen-bond patterns of organic compounds. Acc. Chem. Res. 1990, 23, 120–126. [Google Scholar] [CrossRef]
  25. Mohammad, M.A.; Alhalaweh, A.; Velaga, S.P. Hansen solubility parameter as a tool to predict cocrystal formation. Int. J. Pharm. 2011, 407, 63–71. [Google Scholar] [CrossRef] [PubMed]
  26. Desiraju, G.R. Supramolecular synthons in crystal engineering—A new organic synthesis. Angew. Chem. Int. Ed. Engl. 1995, 34, 2311–2327. [Google Scholar] [CrossRef]
  27. Fábián, L. Cambridge Structural Database Analysis of Molecular Complementarity in Cocrystals. Cryst. Growth Des. 2009, 9, 1436–1443. [Google Scholar] [CrossRef]
  28. Manin, A.N.; Drozd, K.V.; Churakov, A.V.; Perlovich, G.L. Hydrogen Bond Donor/Acceptor Ratios of the Coformers: Do They Really Matter for the Prediction of Molecular Packing in Cocrystals? The Case of Benzamide Derivatives with Dicarboxylic Acids. Cryst. Growth Des. 2018, 18, 5254–5269. [Google Scholar] [CrossRef]
  29. Kumar, S. Pharmaceutical cocrystals: An overview. Indian J. Pharm. Sci. 2018, 79, 858–871. [Google Scholar] [CrossRef]
  30. Surov, A.O.; Ramazanova, A.G.; Voronin, A.P.; Drozd, K.V.; Churakov, A.V.; Perlovich, G.L. Virtual Screening, Structural Analysis, and Formation Thermodynamics of Carbamazepine Cocrystals. Pharmaceutics 2023, 15, 836. [Google Scholar] [CrossRef]
  31. Hansen, C.M. The Three Dimensional Solubility Parameter and Solvent Diffusion Coefficient: Their Importance in Surface Coating Formulation. Ph.D. Dissertation, Danish Technical, Copenhagen, Denmark, 1967. [Google Scholar]
  32. Qiao, N.; Li, M.; Schlindwein, W.; Malek, N.; Davies, A.; Trappitt, G. Pharmaceutical cocrystals: An overview. Int. J. Pharm. 2011, 419, 1–11. [Google Scholar] [CrossRef]
  33. Vishweshwar, P.; McMahon, J.A.; Bis, J.A.; Zaworotko, M.J. Pharmaceutical co-crystals. J. Pharm. Sci. 2006, 95, 499–516. [Google Scholar] [CrossRef]
  34. Kuminek, G.; Cao, F.; Bahia de Oliveira da Rocha, A.; Goncalves Cardoso, S.; Rodriguez-Hornedo, N. Cocrystals to facilitate delivery of poorly soluble compounds beyond-rule-of-5. Adv. Drug Deliv. Rev. 2016, 101, 143–166. [Google Scholar] [CrossRef] [Green Version]
  35. CCDC. Current Structures in the Cambridge Structural Database. Available online: https://www.ccdc.cam.ac.uk/ (accessed on 3 January 2023).
  36. Thakuria, R.; Sarma, B.; Nangia, A. 7.03—Hydrogen Bonding in Molecular Crystals. In Comprehensive Supramolecular Chemistry II; Atwood, J.L., Ed.; Elsevier: Oxford, UK, 2017; pp. 25–48. [Google Scholar]
  37. 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]
  38. Bennett, A.J.; Matzger, A.J. Progress in Predicting Ionic Cocrystal Formation: The Case of Ammonium Nitrate. Chem. Eur. J. 2023, e202300076. [Google Scholar] [CrossRef] [PubMed]
  39. Arunan, E.; Desiraju, G.R.; Klein, R.A.; Sadlej, J.; Scheiner, S.; Alkorta, I.; Clary, D.C.; Crabtree, R.H.; Dannenberg, J.J.; Hobza, P.; et al. Definition of the hydrogen bond (IUPAC Recommendations 2011). Pure Appl. Chem. 2011, 83, 1637–1641. [Google Scholar] [CrossRef]
  40. Etter, M.C. Hydrogen bonds as design elements in organic chemistry. J. Phys. Chem. 1991, 95, 4601–4610. [Google Scholar] [CrossRef]
  41. Cruz-Cabeza, A.J. Acid–base crystalline complexes and the pKa rule. CrystEngComm 2012, 14, 6362–6365. [Google Scholar] [CrossRef]
  42. Goswami, P.K.; Thaimattam, R.; Ramanan, A. Crystal Engineering of Multicomponent Crystal Forms of p-Aminosalicylic Acid with Pyridine Based Coformers. Cryst. Growth Des. 2016, 16, 1268–1281. [Google Scholar] [CrossRef]
  43. Tsume, Y.; Mudie, D.M.; Langguth, P.; Amidon, G.E.; Amidon, G.L. The Biopharmaceutics Classification System: Subclasses for in vivo predictive dissolution (IPD) methodology and IVIVC. Eur. J. Pharm. Sci. 2014, 57, 152–163. [Google Scholar] [CrossRef] [Green Version]
  44. Banik, M.; Gopi, S.P.; Ganguly, S.; Desiraju, G.R. Cocrystal and Salt Forms of Furosemide: Solubility and Diffusion Variations. Cryst. Growth Des. 2016, 16, 5418–5428. [Google Scholar] [CrossRef]
  45. Surov, A.O.; Voronin, A.P.; Manin, A.N.; Manin, N.G.; Kuzmina, L.G.; Churakov, A.V.; Perlovich, G.L. Pharmaceutical cocrystals of diflunisal and diclofenac with theophylline. Mol. Pharm. 2014, 11, 3707–3715. [Google Scholar] [CrossRef]
  46. Nugrahani, I.; Utami, D.; Ibrahim, S.; Nugraha, Y.P.; Uekusa, H. Zwitterionic cocrystal of diclofenac and l-proline: Structure determination, solubility, kinetics of cocrystallization, and stability study. Eur. J. Pharm. Sci. 2018, 117, 168–176. [Google Scholar] [CrossRef]
  47. Aitipamula, S.; Wong, A.B.H.; Chow, P.S.; Tan, R.B.H. Cocrystallization with flufenamic acid: Comparison of physicochemical properties of two pharmaceutical cocrystals. CrystEngComm 2014, 16, 5793–5801. [Google Scholar] [CrossRef]
  48. Goyal, P.; Rani, D.; Chadha, R. Crystal Engineering: A Remedy To Tailor the Biopharmaceutical Aspects of Glibenclamide. Cryst. Growth Des. 2017, 18, 105–118. [Google Scholar] [CrossRef]
  49. Mehta, B.K.; Singh, S.S.; Chaturvedi, S.; Wahajuddin, M.; Thakur, T.S. Rational Coformer Selection and the Development of New Crystalline Multicomponent Forms of Resveratrol with Enhanced Water Solubility. Cryst. Growth Des. 2018, 18, 1581–1592. [Google Scholar] [CrossRef]
  50. Yang, C.; Guo, W.; Lin, Y.; Lin, Q.; Wang, J.; Wang, J.; Zeng, Y. Experimental and DFT simulation study of a novel felodipine cocrystal: Characterization, dissolving properties and thermal decomposition kinetics. J. Pharm. Biomed. Anal. 2018, 154, 198–206. [Google Scholar] [CrossRef]
  51. Kerr, H.E.; Softley, L.K.; Suresh, K.; Nangia, A.; Hodgkinson, P.; Evans, I.R. A furosemide–isonicotinamide cocrystal: An investigation of properties and extensive structural disorder. CrystEngComm 2015, 17, 6707–6715. [Google Scholar] [CrossRef] [Green Version]
  52. Khare, S.G.; Jena, S.K.; Sangamwar, A.T.; Khullar, S.; Mandal, S.K. Multicomponent Pharmaceutical Adducts of α-Eprosartan: Physicochemical Properties and Pharmacokinetic Study. Cryst. Growth Des. 2017, 17, 1589–1599. [Google Scholar] [CrossRef]
  53. Sarkar, A.; Rohani, S. Cocrystals of acyclovir with promising physicochemical properties. J. Pharm. Sci. 2015, 104, 98–105. [Google Scholar] [CrossRef] [PubMed]
  54. Sanphui, P.; Devi, V.K.; Clara, D.; Malviya, N.; Ganguly, S.; Desiraju, G.R. Cocrystals of Hydrochlorothiazide: Solubility and Diffusion/Permeability Enhancements through Drug-Coformer Interactions. Mol. Pharm. 2015, 12, 1615–1622. [Google Scholar] [CrossRef]
  55. 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]
  56. Wang, J.-R.; Ye, C.; Zhu, B.; Zhou, C.; Mei, X. Pharmaceutical cocrystals of the anti-tuberculosis drug pyrazinamide with dicarboxylic and tricarboxylic acids. CrystEngComm 2015, 17, 747–752. [Google Scholar] [CrossRef]
  57. Chattoraj, S.; Shi, L.; Sun, C.C. Understanding the relationship between crystal structure, plasticity and compaction behaviour of theophylline, methyl gallate, and their 1: 1 co-crystal. CrystEngComm 2010, 12, 2466–2472. [Google Scholar] [CrossRef]
  58. Karki, S.; Friščić, T.; Fabian, L.; Laity, P.R.; Day, G.M.; Jones, W. Improving mechanical properties of crystalline solids by cocrystal formation: New compressible forms of paracetamol. Adv. Mater. 2009, 21, 3905–3909. [Google Scholar] [CrossRef]
  59. Sun, C.C.; Hou, H. Improving mechanical properties of caffeine and methyl gallate crystals by cocrystallization. Cryst. Growth Des. 2008, 8, 1575–1579. [Google Scholar] [CrossRef] [Green Version]
  60. Ahmed, H.; Shimpi, M.R.; Velaga, S.P. Relationship between mechanical properties and crystal structure in cocrystals and salt of paracetamol. Drug. Dev. Ind. Pharm. 2017, 43, 89–97. [Google Scholar] [CrossRef] [PubMed]
  61. Rao Khandavilli, U.B.; Bhogala, B.R.; Maguire, A.R.; Lawrence, S.E. Symmetry assisted tuning of bending and brittle multi-component forms of probenecid. Chem. Commun. 2017, 53, 3381–3384. [Google Scholar] [CrossRef] [PubMed]
  62. Nath, N.K.; Hazarika, M.; Gupta, P.; Ray, N.R.; Paul, A.K.; Nauha, E. Plastically bendable crystals of probenecid and its cocrystal with 4,4′-Bipyridine. J. Mol. Struct. 2018, 1160, 20–25. [Google Scholar] [CrossRef]
  63. Chow, S.F.; Chen, M.; Shi, L.; Chow, A.H.; Sun, C.C. Simultaneously improving the mechanical properties, dissolution performance, and hygroscopicity of ibuprofen and flurbiprofen by cocrystallization with nicotinamide. Pharm. Res. 2012, 29, 1854–1865. [Google Scholar] [CrossRef]
  64. Maeno, Y.; Fukami, T.; Kawahata, M.; Yamaguchi, K.; Tagami, T.; Ozeki, T.; Suzuki, T.; Tomono, K. Novel pharmaceutical cocrystal consisting of paracetamol and trimethylglycine, a new promising cocrystal former. Int. J. Pharm. 2014, 473, 179–186. [Google Scholar] [CrossRef]
  65. Sanphui, P.; Mishra, M.K.; Ramamurty, U.; Desiraju, G.R. Tuning mechanical properties of pharmaceutical crystals with multicomponent crystals: Voriconazole as a case study. Mol. Pharm. 2015, 12, 889–897. [Google Scholar] [CrossRef]
  66. Thipparaboina, R.; Kumar, D.; Mittapalli, S.; Balasubramanian, S.; Nangia, A.; Shastri, N.R. Ionic, Neutral, and Hybrid Acid–Base Crystalline Adducts of Lamotrigine with Improved Pharmaceutical Performance. Cryst. Growth Des. 2015, 15, 5816–5826. [Google Scholar] [CrossRef]
  67. Hiendrawan, S.; Veriansyah, B.; Widjojokusumo, E.; Soewandhi, S.N.; Wikarsa, S.; Tjandrawinata, R.R. Physicochemical and mechanical properties of paracetamol cocrystal with 5-nitroisophthalic acid. Int. J. Pharm. 2016, 497, 106–113. [Google Scholar] [CrossRef] [PubMed]
  68. Serrano, D.R.; O’Connell, P.; Paluch, K.J.; Walsh, D.; Healy, A.M. Cocrystal habit engineering to improve drug dissolution and alter derived powder properties. J. Pharm. Pharm. 2016, 68, 665–677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Seo, J.-W.; Hwang, K.-M.; Lee, S.-H.; Kim, D.-W.; Park, E.-S. Preparation and characterization of adefovir dipivoxil–stearic acid cocrystal with enhanced physicochemical properties. Pharm. Dev. Technol. 2017, 23, 890–899. [Google Scholar] [CrossRef]
  70. Haneef, J.; Amir, M.; Sheikh, N.A.; Chadha, R. Mitigating Drug Stability Challenges Through Cocrystallization. AAPS PharmSciTech 2023, 24, 62. [Google Scholar] [CrossRef] [PubMed]
  71. Trask, A.V.; Motherwell, W.S.; Jones, W. Pharmaceutical cocrystallization: Engineering a remedy for caffeine hydration. Cryst. Growth Des. 2005, 5, 1013–1021. [Google Scholar] [CrossRef]
  72. Mittapalli, S.; Bolla, G.; Perumalla, S.; Nangia, A. Can we exchange water in a hydrate structure: A case study of etoricoxib. CrystEngComm 2016, 18, 2825–2829. [Google Scholar] [CrossRef]
  73. Vangala, V.R.; Chow, P.S.; Tan, R.B.H. Characterization, physicochemical and photo-stability of a co-crystal involving an antibioticdrug, nitrofurantoin, and 4-hydroxybenzoic acid. CrystEngComm 2011, 13, 759–762. [Google Scholar] [CrossRef]
  74. Babu, N.J.; Sanphui, P.; Nangia, A. Crystal engineering of stable temozolomide cocrystals. Chem. Asian J. 2012, 7, 2274–2285. [Google Scholar] [CrossRef]
  75. Geng, N.; Chen, J.-M.; Li, Z.-J.; Jiang, L.; Lu, T.-B. Approach of Cocrystallization to Improve the Solubility and Photostability of Tranilast. Cryst. Growth Des. 2013, 13, 3546–3553. [Google Scholar] [CrossRef]
  76. Jung, S.; Choi, I.; Kim, I. Liquid-Assisted Grinding to Prepare a Cocrystal of Adefovir Dipivoxil Thermodynamically Less Stable than Its Neat Phase. Crystals 2015, 5, 583–591. [Google Scholar] [CrossRef] [Green Version]
  77. Nechipadappu, S.K.; Ramachandran, J.; Shivalingegowda, N.; Lokanath, N.K.; Trivedi, D.R. Synthesis of cocrystals/salts of flucytosine: Structure and stability. New J. Chem. 2018, 42, 5433–5446. [Google Scholar] [CrossRef]
  78. Dai, X.-L.; Li, S.; Chen, J.-M.; Lu, T.-B. Improving the Membrane Permeability of 5-Fluorouracil via Cocrystallization. Cryst. Growth Des. 2016, 16, 4430–4438. [Google Scholar] [CrossRef]
  79. Do Amaral, L.H.; do Carmo, F.A.; Amaro, M.I.; de Sousa, V.P.; da Silva, L.; de Almeida, G.S.; Rodrigues, C.R.; Healy, A.M.; Cabral, L.M. Development and Characterization of Dapsone Cocrystal Prepared by Scalable Production Methods. AAPS PharmSciTech 2018, 19, 2687–2699. [Google Scholar] [CrossRef] [PubMed]
  80. Balimane, P.V.; Chong, S. Evaluation of Permeability and P-glycoprotein Interactions: Industry Outlook. In Biopharmaceutics Applications in Drug Development; Krishna, R., Yu, L., Eds.; Springer: Boston, MA, USA, 2008; pp. 101–138. [Google Scholar]
  81. Khatioda, R.; Bora, P.; Sarma, B. Trimorphic Ethenzamide Cocrystal: In Vitro Solubility and Membrane Efflux Studies. Cryst. Growth Des. 2018, 18, 4637–4645. [Google Scholar] [CrossRef]
  82. Bommaka, M.K.; Mannava, M.K.C.; Suresh, K.; Gunnam, A.; Nangia, A. Entacapone: Improving Aqueous Solubility, Diffusion Permeability, and Cocrystal Stability with Theophylline. Cryst. Growth Des. 2018, 18, 6061–6069. [Google Scholar] [CrossRef]
  83. Surov, A.O.; Volkova, T.V.; Churakov, A.V.; Proshin, A.N.; Terekhova, I.V.; Perlovich, G.L. Cocrystal formation, crystal structure, solubility and permeability studies for novel 1,2,4-thiadiazole derivative as a potent neuroprotector. Eur. J. Pharm. Sci. 2017, 109, 31–39. [Google Scholar] [CrossRef]
  84. Gopi, S.P.; Banik, M.; Desiraju, G.R. New Cocrystals of Hydrochlorothiazide: Optimizing Solubility and Membrane Diffusivity. Cryst. Growth Des. 2016, 17, 308–316. [Google Scholar] [CrossRef]
  85. Sanphui, P.; Babu, N.J.; Nangia, A. Temozolomide Cocrystals with Carboxamide Coformers. Cryst. Growth Des. 2013, 13, 2208–2219. [Google Scholar] [CrossRef]
  86. Cho, M.-Y.; Kim, P.; Kim, G.-Y.; Lee, J.-Y.; Song, K.-H.; Lee, M.-J.; Yoon, W.; Yun, H.; Choi, G.J. Preparation and Characterization of Aripiprazole Cocrystals with Coformers of Multihydroxybenzene Compounds. Cryst. Growth Des. 2017, 17, 6641–6652. [Google Scholar] [CrossRef]
  87. Deka, P.; Gogoi, D.; Althubeiti, K.; Rao, D.R.; Thakuria, R. Mechanosynthesis, Characterization, and Physicochemical Property Investigation of a Favipiravir Cocrystal with Theophylline and GRAS Coformers. Cryst. Growth Des. 2021, 21, 4417–4425. [Google Scholar] [CrossRef]
  88. Sathisaran, I.; Dalvi, S.V. Crystal Engineering of Curcumin with Salicylic Acid and Hydroxyquinol as Coformers. Cryst. Growth Des. 2017, 17, 3974–3988. [Google Scholar] [CrossRef]
  89. Gołdyn, M.R.; Larowska, D.; Bartoszak-Adamska, E. Novel Purine Alkaloid Cocrystals with Trimesic and Hemimellitic Acids as Coformers: Synthetic Approach and Supramolecular Analysis. Cryst. Growth Des. 2020, 21, 396–413. [Google Scholar] [CrossRef]
  90. 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]
  91. Hrinova, E.; Skorepova, E.; Cerna, I.; Kralovicova, J.; Kozlik, P.; Krizek, T.; Rousarova, J.; Rysanek, P.; Sima, M.; Slanar, O.; et al. Explaining dissolution properties of rivaroxaban cocrystals. Int. J. Pharm. 2022, 622, 121854. [Google Scholar] [CrossRef] [PubMed]
  92. Meng, Y.; Tan, F.; Yao, J.; Cui, Y.; Feng, Y.; Li, Z.; Wang, Y.; Yang, Y.; Gong, W.; Yang, M.; et al. Preparation, characterization, and pharmacokinetics of rivaroxaban cocrystals with enhanced in vitro and in vivo properties in beagle dogs. Int. J. Pharm. X 2022, 4, 100119. [Google Scholar] [CrossRef]
  93. Kale, D.P.; Puri, V.; Kumar, A.; Kumar, N.; Bansal, A.K. The Role of Cocrystallization-Mediated Altered Crystallographic Properties on the Tabletability of Rivaroxaban and Malonic Acid. Pharmaceutics 2020, 12, 546. [Google Scholar] [CrossRef] [PubMed]
  94. Kale, D.P.; Ugale, B.; Nagaraja, C.M.; Dubey, G.; Bharatam, P.V.; Bansal, A.K. Molecular Basis of Water Sorption Behavior of Rivaroxaban-Malonic Acid Cocrystal. Mol. Pharm. 2019, 16, 2980–2991. [Google Scholar] [CrossRef] [PubMed]
  95. Kuang, W.j.; Ji, S.c.; Xu, S.m.; Lan, P.; Liao, A.p.; Zhou, J.y.; Zhang, J.y. Thermodynamic and Crystallization of Lamotrigine Cocrystal. Cryst. Growth Des. 2019, 19, 6603–6610. [Google Scholar] [CrossRef]
  96. Lekšić, E.; Pavlović, G.; Meštrović, E. Cocrystals of Lamotrigine Based on Coformers Involving Carbonyl Group Discovered by Hot-Stage Microscopy and DSC Screening. Cryst. Growth Des. 2012, 12, 1847–1858. [Google Scholar] [CrossRef]
  97. Kuang, W.; Ji, S.; Wang, X.; Zhang, J.; Lan, P. Relationship between crystal structures and physicochemical properties of lamotrigine cocrystal. Powder Technol. 2021, 380, 18–25. [Google Scholar] [CrossRef]
  98. Kaur, R.; Cavanagh, K.L.; Rodriguez-Hornedo, N.; Matzger, A.J. Multidrug Cocrystal of Anticonvulsants: Influence of Strong Intermolecular Interactions on Physiochemical Properties. Cryst. Growth Des. 2017, 17, 5012–5016. [Google Scholar] [CrossRef] [PubMed]
  99. Du, S.; Wang, Y.; Wu, S.; Yu, B.; Shi, P.; Bian, L.; Zhang, D.; Hou, J.; Wang, J.; Gong, J. Two novel cocrystals of lamotrigine with isomeric bipyridines and in situ monitoring of the cocrystallization. Eur. J. Pharm. Sci. 2017, 110, 19–25. [Google Scholar] [CrossRef] [PubMed]
  100. Chadha, R.; Saini, A.; Arora, P.; Jain, D.S.; Dasgupta, A.; Guru Row, T.N. Multicomponent solids of lamotrigine with some selected coformers and their characterization by thermoanalytical, spectroscopic and X-ray diffraction methods. CrystEngComm 2011, 13, 6271–6284. [Google Scholar] [CrossRef]
  101. Zalte, A.G.; Saudagar, R.B. Preparation and Characterization of Etodolac Co-Crystals Using 32 Full Factorial Design. Res. J. Pharm. Technol. 2018, 11, 3781–3786. [Google Scholar] [CrossRef]
  102. Ahirrao, S.P.; Sonawane, M.P.; Bhambere, D.S.; Udavant, P.B.; Ahire, E.D.; Kanade, R.; kuber, D. Cocrystal Formulation: A Novel Approach to Enhance Solubility and Dissolution of Etodolac. Biosci. Biotechnol. Res. Asia 2022, 19, 111–119. [Google Scholar] [CrossRef]
  103. Zachariah, S.B.; Borges, E.F.; Varghese, R.; Cruz, A.R.; Ross, G.S. Positive response to oral divalproex sodium (Depakote) in patients with spasticity and pain. Am. J. Med. Sci. 1994, 308, 38–40. [Google Scholar] [CrossRef]
  104. McGraw, D. Therapeutic drug monitoring with valproate–Why product selection is an important factor. Ment. Health Clin. 2014, 4, 31–34. [Google Scholar] [CrossRef]
  105. Kavanagh, O.N.; Croker, D.M.; Walker, G.M.; Zaworotko, M.J. Pharmaceutical cocrystals: From serendipity to design to application. Drug Discov. Today 2019, 24, 796–804. [Google Scholar] [CrossRef] [Green Version]
  106. Shaikh, R.; Singh, R.; Walker, G.M.; Croker, D.M. Pharmaceutical Cocrystal Drug Products: An Outlook on Product Development. Trends Pharm. Sci 2018, 39, 1033–1048. [Google Scholar] [CrossRef]
  107. USFDA. Highlights of Prescribing Information Entresto. Available online: https://www.novartis.com/us-en/sites/novartis_us/files/entresto.pdf (accessed on 23 November 2022).
  108. Astellas. Launch of Suglat. Available online: https://www.astellas.com/en/system/files/news/2018-06/140417_1_Eg.pdf (accessed on 20 October 2022).
  109. European Medicines Agency. ANNEX I, Summary of Product Characteristics Steglatro. Available online: https://www.ema.europa.eu/en/documents/product-information/steglatro-epar-product-information_en.pdf (accessed on 20 October 2022).
  110. USFDA. Highlights of Prescribing Information Steglatro. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2017/209803s000lbl.pdf (accessed on 20 October 2022).
  111. Estevez, K.L.A.R.R. Escitalopram. Available online: https://www.ncbi.nlm.nih.gov/books/NBK557734/ (accessed on 20 October 2022).
  112. USFDA. Highlights of Prescribing Information Lexapro. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2017/021323s047lbl.pdf (accessed on 20 October 2022).
  113. USFDA. Drug Approval Package. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/nda/99/020793_000_CafcitTOC.cfm (accessed on 20 October 2022).
  114. Evaluation and Licensing Division, Pharmaceutical and Food Safety Bureau; Ministry of Health, Labour and Welfare. Japanese Review Report. Available online: https://www.pmda.go.jp/files/000213963.pdf (accessed on 20 October 2022).
  115. Bhatt, P.M.; Azim, Y.; Thakur, T.S.; Desiraju, G.R. Co-crystals of the anti-HIV drugs lamivudine and zidovudine. Cryst. Growth Des. 2009, 9, 951–957. [Google Scholar] [CrossRef]
  116. European Medicines Agency. European Public Assessment Report (EPAR). Available online: https://www.ema.europa.eu/en/medicines/human/EPAR/lamivudine-zidovudine-teva#product-information-section (accessed on 20 October 2022).
  117. European Medicines Agency. ANNEX I, Summary of Product Characteristics Lamivudine/Zidovudine. Available online: https://www.ema.europa.eu/en/documents/product-information/lamivudine/zidovudine-teva-epar-product-information_en.pdf (accessed on 21 October 2022).
  118. Alvani, A.; Shayanfar, A. Solution Stability of Pharmaceutical Cocrystals. Cryst. Growth Des. 2022, 22, 6323–6337. [Google Scholar] [CrossRef]
  119. San Martín, O.; Llombart, B.; Carretero Hernandez, G.; Flórez Menéndez, Á.; Botella-Estrada, R.; Herrera Ceballos, E.; Puig, S. Sonidegib in the Treatment of Locally Advanced Basal Cell Carcinoma. Actas Dermo-Sifiliográficas 2021, 112, 295–301. [Google Scholar] [CrossRef]
  120. USFDA. Highlights of Prescribing Information of ODOMZO®. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2019/205266s006lbl.pdf (accessed on 22 October 2022).
  121. Diaz-Villamarin, X.; Pinar-Morales, R.; Barrero-Hernandez, F.J.; Antunez-Rodriguez, A.; Cabeza-Barrera, J.; Moron-Romero, R. Pharmacogenetics of siponimod: A systematic review. Biomed. Pharm. 2022, 153, 113536. [Google Scholar] [CrossRef]
  122. O’Sullivan, A.; Long, B.; Verma, V.; Ryan, K.M.; Padrela, L. Solid-state and particle size control of pharmaceutical cocrystals using atomization-based techniques. Int. J. Pharm. 2022, 621, 121798. [Google Scholar] [CrossRef] [PubMed]
  123. USFDA. Highlights of Prescribing Information of MAYZENT®. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2022/209884Orig1s011lbl.pdf (accessed on 22 October 2022).
  124. USFDA. Opioid Analgesic REMS Program. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2021/213426s000lbl.pdf (accessed on 23 October 2022).
  125. Putra, O.D.; Yoshida, T.; Umeda, D.; Higashi, K.; Uekusa, H.; Yonemochi, E. Crystal Structure Determination of Dimenhydrinate after More than 60 Years: Solving Salt–Cocrystal Ambiguity via Solid-State Characterizations and Solubility Study. Cryst. Growth Des. 2016, 16, 5223–5229. [Google Scholar] [CrossRef]
  126. USFDA. Orange Book: Approved Drug Products with Therapeutic Equivalence Evaluations. Available online: https://www.accessdata.fda.gov/scripts/cder/ob/results_product.cfm?Appl_Type=A&Appl_No=085166#13840 (accessed on 23 October 2022).
  127. Maytal Piran, Teva Pharmaceutical Industries Ltd. Creatively Using Co-Crystals to Produce Alternative Generic Products. Available online: https://www.teva-api.com/knowledge-center/using-co-crystals-to-produce-alternative-generic-products/ (accessed on 24 October 2022).
  128. Banerjee, M.; Nimkar, K.; Naik, S.; Patravale, V. Unlocking the potential of drug-drug cocrystalsA comprehensive review. J. Control Release 2022, 348, 456–469. [Google Scholar] [CrossRef]
  129. Kumar, A.; Kumar, S.; Nanda, A. A Review about Regulatory Status and Recent Patents of Pharmaceutical Co-Crystals. Adv. Pharm. Bull. 2018, 8, 355–363. [Google Scholar] [CrossRef]
  130. Videla, S.; Gascón, N.; Plata-Salamán, C. Commentary on the “Co-Crystal of Tramadol-Celecoxib in Patients with Moderate to Severe Acute Post-Surgical Oral Pain: A Dose-Finding, Randomised, Double-Blind, Placebo-and Active-Controlled, Multicentre Phase II Trial”. Implications for Cardiovascular Safety. J. Cardiol. Cardiovasc. Sci. 2019, 3. [Google Scholar] [CrossRef] [Green Version]
  131. Merlos, M.; Portillo-Salido, E.; Brenchat, A.; Aubel, B.; Buxens, J.; Fisas, A.; Codony, X.; Romero, L.; Zamanillo, D.; Vela, J.M. Administration of a co-crystal of tramadol and celecoxib in a 1: 1 molecular ratio produces synergistic antinociceptive effects in a postoperative pain model in rats. Eur. J. Pharmacol. 2018, 833, 370–378. [Google Scholar] [CrossRef]
  132. U.S. National Library of Medicine. Co-Crystal E-58425 vs. Tramadol and Celecoxib for Moderate to Severe Acute Pain after Bunionectomy. Phase III Clinical Trial. Available online: https://clinicaltrials.gov/ct2/show/NCT03108482 (accessed on 24 October 2022).
  133. Kimoto, K.; Yamamoto, M.; Karashima, M.; Hohokabe, M.; Takeda, J.; Yamamoto, K.; Ikeda, Y. Pharmaceutical Cocrystal Development of TAK-020 with Enhanced Oral Absorption. Crystals 2020, 10, 211. [Google Scholar] [CrossRef] [Green Version]
  134. U.S. National Library of Medicine. In TAK-020 Relative Bioavailability and Food Effect Study in Healthy Participants. Available online: https://clinicaltrials.gov/ct2/show/NCT02723201 (accessed on 24 October 2022).
  135. Devarakonda, S.N.; Vyas, K.; Bommareddy, S.R.; Padi, P.R.; Raghupathy, B. ARIPIPRAZOLE CO-CRYSTALS. US 2009/0054455, 2009. [Google Scholar]
  136. National Library of Medicine-National Center for Biotechnology Information. Compound Summary of Fumaric acid. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/Fumaric-acid (accessed on 27 October 2022).
  137. Drugbank. Fumaric Acid. Available online: https://go.drugbank.com/drugs/DB01677 (accessed on 26 October 2022).
  138. Yalkowsky, S.H.; He, Y.; Jain, P. Handbook of Aqueous Solubility Data, 2nd ed.; CRC Press, Taylor & Francis Group: Boca Raton, FL, USA, 2010. [Google Scholar]
  139. Roa Engel, C.A.; Straathof, A.J.; Zijlmans, T.W.; van Gulik, W.M.; van der Wielen, L.A. Fumaric acid production by fermentation. Appl. Microbiol. Biotechnol. 2008, 78, 379–389. [Google Scholar] [CrossRef] [Green Version]
  140. National Library of Medicine-National Center for Biotechnology Information. Compound Summary of Oxalic Acid. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/Oxalic-acid#section=Related-Compounds (accessed on 27 October 2022).
  141. Palmieri, F.; Estoppey, A.; House, G.L.; Lohberger, A.; Bindschedler, S.; Chain, P.S.; Junier, P. Oxalic acid, a molecule at the crossroads of bacterial-fungal interactions. Adv. Appl. Microbiol. 2019, 106, 49–77. [Google Scholar] [CrossRef]
  142. Saxena, R.K.; Saran, S.; Isar, J.; Kaushik, R. Production and Applications of Succinic Acid. In Current Developments in Biotechnology and Bioengineering; Elsevier: Amsterdam, The Netherlands, 2017; pp. 601–630. [Google Scholar]
  143. Yan, Y.; Wang, L.; Si, Z.; Zhang, X.; Yuan, W. A novel cocrystal of metformin hydrochloride with citric acid: Systematic synthesis and computational simulation. Eur. J. Pharm. Biopharm. 2022, 179, 37–46. [Google Scholar] [CrossRef] [PubMed]
  144. National Library of Medicine-National Center for Biotechnology Information. Compound Summary of Citric Acid. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/Citric-Acid (accessed on 27 October 2022).
  145. Sebastian, J.; Osorio-Gonzalez, C.; Rouissi, T.; Hegde, K.; Brar, S.K. Bioderived fumaric acid for sustainable production of key active pharmaceutical ingredients: Dimethyl fumarate and Monomethyl fumarate. Process Biochem. 2022, 120, 35–40. [Google Scholar] [CrossRef]
  146. Borodi, G.; Turza, A.; Onija, O.; Bende, A. Succinic, fumaric, adipic and oxalic acid cocrystals of promethazine hydrochloride. Acta Cryst. C Struct. Chem. 2019, 75, 107–119. [Google Scholar] [CrossRef] [PubMed]
  147. Sathisaran, I.; Dalvi, S.V. Engineering Cocrystals of PoorlyWater-Soluble Drugs to Enhance Dissolution in Aqueous Medium. Pharmaceutics 2018, 10, 108. [Google Scholar] [CrossRef] [Green Version]
  148. Shan, N.; Zaworotko, M.J. The role of cocrystals in pharmaceutical science. Drug Discov. Today 2008, 13, 440–446. [Google Scholar] [CrossRef]
  149. Guo, M.; Sun, X.; Chen, J.; Cai, T. Pharmaceutical cocrystals: A review of preparations, physicochemical properties and applications. Acta Pharm. Sin. B 2021, 11, 2537–2564. [Google Scholar] [CrossRef]
  150. Liu, L.; Zou, D.; Zhang, Y.; Zhang, Q.; Feng, Y.; Guo, Y.; Liu, Y.; Zhang, X.; Cheng, G.; Wang, C.; et al. Pharmaceutical salts/cocrystals of enoxacin with dicarboxylic acids: Enhancing in vitro antibacterial activity of enoxacin by improving the solubility and permeability. Eur. J. Pharm. Biopharm. 2020, 154, 62–73. [Google Scholar] [CrossRef]
  151. Childs, S.L.; Chyall, L.J.; Dunlap, J.T.; Smolenskaya, V.N.; Stahly, B.C.; Stahly, G.P. Crystal engineering approach to forming cocrystals of amine hydrochlorides with organic acids. Molecular complexes of fluoxetine hydrochloride with benzoic, succinic, and fumaric acids. J. Am. Chem. Soc. 2004, 126, 13335–13342. [Google Scholar] [CrossRef] [Green Version]
  152. Yogheshwari, P.; Sridhar, B.; Anitha, K. Experimental and theoretical studies on bis (6-nitroquinoline) fumaric acid co-crystal. J. Mol. Struct. 2022, 1249, 131561. [Google Scholar] [CrossRef]
  153. Pandey, N.; Ghosh, A. An outlook on permeability escalation through cocrystallization for developing pharmaceuticals with improved biopharmaceutical properties. J. Drug Deliv. Sci. Technol. 2022, 76, 103757. [Google Scholar] [CrossRef]
  154. Mohamed, M.P.; Sudha, S.; Jayaprakash, P.; Vinitha, G.; Nageshwari, M.; Sangeetha, P.; Kumari, C.R.T.; Caroline, M.L. Growth and characterization of L-histidinium fumarate fumaric acid monohydrate single crystal: A promising second and third order nonlinear optical material. Chin. J. Phys. 2019, 60, 581–597. [Google Scholar] [CrossRef]
  155. Prakash, M.; Geetha, D.; Caroline, M. Growth and Characterization of Nonlinear Optics (NLO) Active L-Phenylalanine Fumaric Acid (LPFA) Single Crystal. Mater. Manuf. Process. 2011, 27, 519–522. [Google Scholar] [CrossRef]
  156. Sawatdee, S.; Atipairin, A.; Rakkummerd, S.; Suriyaphol, O.; Harding, D.J.; Muenraya, P.; Harding, P. Preparation and physicochemical characterization of sildenafil cocrystals. J. Adv. Pharm. Technol. Res. 2021, 12, 408–419. [Google Scholar] [CrossRef]
  157. Kamble, R.N.; Bothiraja, C.; Mehta, P.P.; Varghese, V. Synthesis, solid state characterization and antifungal activity of ketoconazole cocrystals. J. Pharm. Investig. 2017, 48, 541–549. [Google Scholar] [CrossRef]
  158. Maheshwari, C.; André, V.; Reddy, S.; Roy, L.; Duarte, T.; Rodríguez-Hornedo, N. Tailoring aqueous solubility of a highly soluble compound via cocrystallization: Effect of coformer ionization, pH max and solute–solvent interactions. CrystEngComm 2012, 14, 4801–4811. [Google Scholar] [CrossRef]
  159. Fernandes, R.P.; do Nascimento, A.L.C.S.; Carvalho, A.C.S.; Teixeira, J.A.; Ionashiro, M.; Caires, F.J. Mechanochemical synthesis, characterization, and thermal behavior of meloxicam cocrystals with salicylic acid, fumaric acid, and malic acid. J. Therm. Anal. Calorim. 2019, 138, 765–777. [Google Scholar] [CrossRef]
  160. Li, L.; Yin, X.H.; Diao, K.S. Improving the solubility and bioavailability of anti-hepatitis B drug PEC via PEC-fumaric acid cocrystal. RSC Adv. 2020, 10, 36125–36134. [Google Scholar] [CrossRef] [PubMed]
  161. Chadha, R. Novel Cocrystals of Glipizide: Green Supramolecular Mechanosynthesis. Arch. Pharm. Pharmacol. Res. 2018, 1, 1–13. [Google Scholar] [CrossRef]
  162. Bruni, G.; Maietta, M.; Maggi, L.; Mustarelli, P.; Ferrara, C.; Berbenni, V.; Milanese, C.; Girella, A.; Marini, A. Preparation and physicochemical characterization of acyclovir cocrystals with improved dissolution properties. J. Pharm. Sci. 2013, 102, 4079–4086. [Google Scholar] [CrossRef] [PubMed]
  163. Aitipamula, S.; Wong, A.B.H.; Chow, P.S.; Tan, R.B.H. Pharmaceutical cocrystals of ethenzamide: Structural, solubility and dissolution studies. CrystEngComm 2012, 14, 8515–8524. [Google Scholar] [CrossRef]
  164. Dayo Owoyemi, B.C.; da Silva, C.C.P.; Souza, M.S.; Diniz, L.F.; Ellena, J.; Carneiro, R.L. Fluconazole: Synthesis and Structural Characterization of Four New Pharmaceutical Cocrystal Forms. Cryst. Growth Des. 2019, 19, 648–657. [Google Scholar] [CrossRef]
  165. Gadade, D.D.; Pekamwar, S.S.; Shirsat, M.D. Crystal Engineering of Antiviral Agent Efavirenz for Solubility Enhancement. J. Drug Deliv. Ther. 2018, 8, 86–91. [Google Scholar] [CrossRef] [Green Version]
  166. Chen, Y.; Li, L.; Yao, J.; Ma, Y.-Y.; Chen, J.-M.; Lu, T.-B. Improving the Solubility and Bioavailability of Apixaban via Apixaban–Oxalic Acid Cocrystal. Cryst. Growth Des. 2016, 16, 2923–2930. [Google Scholar] [CrossRef]
  167. Kanakaraju, K.; Lavanya, V.; Nangia, A. Temozolomide Cocrystals Exhibit Drug Sensitivity in Glioblastoma Cells. Proc. Natl. Acad. Sci. India Sect. A Phys. Sci. 2014, 84, 321–330. [Google Scholar] [CrossRef]
  168. Chen, J.Y.; Wu, H.; Guo, C.Y.; Zhu, B.; Ren, G.B. Enhancing the solubility of natural compound xanthotoxin by modulating stability via cocrystallization engineering. Int. J. Pharm. 2019, 572, 118776. [Google Scholar] [CrossRef]
  169. Jindal, A.; Singh, R.; Tomar, S.; Dureja, J.; Karan, M.; Chadha, R. Engineering a Remedy to Modulate and Optimize Biopharmaceutical Properties of Rebamipide by Synthesizing New Cocrystal: In Silico and Experimental Studies. Pharm. Res. 2021, 38, 2129–2145. [Google Scholar] [CrossRef]
  170. Alatas, F.; Ratih, H.; Soewandhi, S.N. Enhancement of solubility and dissolution rate of telmisartan by telmisartan-oxalic acid co-crystal formation. Int. J. Pharm. Pharm. Sci. 2015, 7, 423–426, ISSN- 0975-1491. [Google Scholar]
  171. Aher, S.; Dhumal, R.; Mahadik, K.; Ketolainen, J.; Paradkar, A. Effect of cocrystallization techniques on compressional properties of caffeine/oxalic acid 2:1 cocrystal. Pharm. Dev. Technol. 2011, 18, 55–60. [Google Scholar] [CrossRef]
  172. Amin, S.; Sholihah, R.t.; Megantara, S.; Budiman, A. Synthesis of Glibenclamide-Oxalic Acid Cocrystal using ThermalSolvent-Free Method. Int. J. Pharm. Qual. Assur. 2020, 11, 404–408. [Google Scholar] [CrossRef]
  173. Tretter, L.; Patocs, A.; Chinopoulos, C. Succinate, an intermediate in metabolism, signal transduction, ROS, hypoxia, and tumorigenesis. Biochim. Biophys. Acta 2016, 1857, 1086–1101. [Google Scholar] [CrossRef] [PubMed]
  174. Ninawe, A.; Kumar, A.; Mohadikar, P.; Shende, D.; Wasewar, K. Separation of succinic acid from aqueous phase using nontoxic solvents. Chem. Data Collect. 2022, 39, 100866. [Google Scholar] [CrossRef]
  175. Fuessl, A.; Yamamoto, M.; Schneller, A. Opportunities in Bio-Based Building Blocks for Thermoplastic Polymers. Ref. Modul. Mater. Sci. Mater. Eng. 2016, 1–25. [Google Scholar] [CrossRef]
  176. Marketsandmarkets. Succinic Acid Market. Available online: https://www.marketsandmarkets.com/Market-Reports/succinic-acid-market-402.html (accessed on 24 October 2022).
  177. Alhalaweh, A.; George, S.; Boström, D.; Velaga, S.P. 1:1 and 2:1 Urea−Succinic Acid Cocrystals: Structural Diversity, Solution Chemistry, and Thermodynamic Stability. Cryst. Growth Des. 2010, 10, 4847–4855. [Google Scholar] [CrossRef]
  178. Ober, C.A.; Gupta, R.B. Formation of itraconazole–succinic acid cocrystals by gas antisolvent cocrystallization. Aaps Pharmscitech 2012, 13, 1396–1406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Fitriani, L.; Fitriandi, A.D.; Hasanah, U.; Zaini, E. Nano-Cocrystals of Piperine-Succinic Acid: Physicochemical Characterization and Dissolution Rate Studies. ChemistrySelect 2022, 7, e202104196. [Google Scholar] [CrossRef]
  180. Lemmerer, A.; Bernstein, J.; Kahlenberg, V. One-pot covalent and supramolecular synthesis of pharmaceutical co-crystals using the API isoniazid: A potential supramolecular reagent. CrystEngComm 2010, 12, 2856–2864. [Google Scholar] [CrossRef]
  181. Noonan, T.J.; Chibale, K.; Bourne, S.A.; Caira, M.R. A preformulation co-crystal screening case study: Polymorphic co-crystals of an imidazopyridazine antimalarial drug lead with the coformer succinic acid. J. Mol. Struct. 2020, 1204, 127561. [Google Scholar] [CrossRef]
  182. Rama, V.; Vidavulur, S.; Tadikonda, P.V.; Rajana, N.; Mittapalli, S. Novel cocrystals of brexpiprazole with improved solubility. J. Cryst. Growth 2020, 551, 125910. [Google Scholar] [CrossRef]
  183. Butreddy, A.; Almutairi, M.; Komanduri, N.; Bandari, S.; Zhang, F.; Repka, M.A. Multicomponent crystalline solid forms of aripiprazole produced via hot melt extrusion techniques: An exploratory study. J. Drug Deliv. Sci. Technol. 2021, 63, 102529. [Google Scholar] [CrossRef]
  184. Cao, F.; Rodriguez-Hornedo, N.; Amidon, G.E. Mechanistic Analysis of Cocrystal Dissolution, Surface pH, and Dissolution Advantage as a Guide for Rational Selection. J. Pharm. Sci. 2019, 108, 243–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Chennuru, R.; Devarapalli, R.; Rengaraj, P.; Srinivas, P.L.; Dey, S.; Reddy, C.M. Improving Solubility of Poorly Soluble Abiraterone Acetate by Cocrystal Design Aided by In Silico Screening. Cryst. Growth Des. 2020, 20, 5018–5030. [Google Scholar] [CrossRef]
  186. Rahman, F.; Winantari, A.N.; Setyawan, D.; Siswandono, S. Comparison Study of Grinding and Slurry Method on Physicochemical Characteristic of AcyclovirSuccinic Acid Cocrystal. Asian J. Pharm. Clin. Res. 2017, 10, 153–158. [Google Scholar] [CrossRef]
  187. Poerwono, H.; Higashiyama, K.; Kubo, H.; Poernomo, A.T.; Suharjono; Sudiana, I.K.; Indrayanto, G.; Brittain, H.G. Citric Acid. In Analytical Profiles of Drug Substances and Excipients; Elsevier: Amsterdam, The Netherlands, 2001; Volume 45, pp. 1–76. [Google Scholar]
  188. Research and Markets-The World’s Largest Market Research Store. Citric Acid Market: Global Industry Trends, Share, Size, Growth, Opportunity and Forecast 2022–2027. Available online: https://www.researchandmarkets.com/reports/5577854/citric-acid-market-global-industry-trends (accessed on 27 October 2022).
  189. Hsu, P.-C.; Lin, H.-L.; Wang, S.-L.; Lin, S.-Y. Solid-state thermal behavior and stability studies of theophylline–citric acid cocrystals prepared by neat cogrinding or thermal treatment. J. Solid State Chem. 2012, 192, 238–245. [Google Scholar] [CrossRef]
  190. Deng, J.H.; Lu, T.B.; Sun, C.C.; Chen, J.M. Dapagliflozin-citric acid cocrystal showing better solid state properties than dapagliflozin. Eur. J. Pharm. Sci. 2017, 104, 255–261. [Google Scholar] [CrossRef]
  191. Salem, A.; Khanfar, E.; Nagy, S.; Szechenyi, A. Cocrystals of tuberculosis antibiotics: Challenges and missed opportunities. Int. J. Pharm. 2022, 623, 121924. [Google Scholar] [CrossRef] [PubMed]
  192. Rachmaniar, R.; Riasari, H.; Fauziah, L.; Kenti; Ferdiansyah, R. The effect of cocrystallization method and citric acid as coformer on water solubility of ethyl p-metoxycinnamate particle. In Proceedings of the 3rd International Conference on Condensed Matter and Applied Physics (Icc-2019), Solo, Indonesia, 5 May 2020. [Google Scholar]
  193. Khan, F.M.; Ahmad, M.; Batool, F. Enhancement of solubility and release profile of simvastatin by co-crystallization with citric acid. Trop. J. Pharm. Res. 2019, 18, 2465–2472. [Google Scholar] [CrossRef]
  194. Dhibar, M.; Chakraborty, S.; Basak, S.; Pattanayak, P.; Chatterjee, T.; Ghosh, B.; Raafat, M.; Abourehab, M.A.S. Critical Analysis and Optimization of Stoichiometric Ratio of Drug-Coformer on Cocrystal Design: Molecular Docking, In Vitro and In Vivo Assessment. Pharmaceuticals 2023, 16, 284. [Google Scholar] [CrossRef]
  195. Buol, X.; Robeyns, K.; Caro Garrido, C.; Tumanov, N.; Collard, L.; Wouters, J.; Leyssens, T. Improving Nefiracetam Dissolution and Solubility Behavior Using a Cocrystallization Approach. Pharmaceutics 2020, 12, 653. [Google Scholar] [CrossRef]
  196. Narala, S.; Ampati, R.; Nanam, R. Solubility enhancement of ritonavir: Cocrystallization. J. Pharm. Res. 2019, 8, 630–637. [Google Scholar] [CrossRef]
  197. Alhalaweh, A.; George, S.; Basavoju, S.; Childs, S.L.; Rizvi, S.A.A.; Velaga, S.P. Pharmaceutical cocrystals of nitrofurantoin: Screening, characterization and crystal structure analysis. CrystEngComm 2012, 14, 5078–5088. [Google Scholar] [CrossRef]
  198. Liu, Q.; Yang, D.; Chen, T.; Zhang, B.; Xing, C.; Zhang, L.; Lu, Y.; Du, G. Insights into the Solubility and Structural Features of Four Praziquantel Cocrystals. Cryst. Growth Des. 2021, 21, 6321–6331. [Google Scholar] [CrossRef]
  199. Intellectual Property India. Filing International Applications for Patent Under the Patent Cooperation Treaty. Available online: https://ipindia.gov.in/writereaddata/images/pdf/pct-filing.pdf, (accessed on 20 October 2022).
Figure 1. Techniques utilized for cocrystal preparation.
Figure 1. Techniques utilized for cocrystal preparation.
Pharmaceutics 15 01161 g001
Figure 2. The most common supramolecular synthons in crystal engineering are (a) carboxylic acid dimer homosynthon, (b) carboxylic acid and pyridine group heterosynthon, (c) amide dimer heterosynthon, (d) amide group heterosynthons, and (e) alcohol and ether group heterosynthons.
Figure 2. The most common supramolecular synthons in crystal engineering are (a) carboxylic acid dimer homosynthon, (b) carboxylic acid and pyridine group heterosynthon, (c) amide dimer heterosynthon, (d) amide group heterosynthons, and (e) alcohol and ether group heterosynthons.
Pharmaceutics 15 01161 g002
Figure 3. Properties of coformer and drug to be considered while designing the cocrystal.
Figure 3. Properties of coformer and drug to be considered while designing the cocrystal.
Pharmaceutics 15 01161 g003
Figure 4. Pharmaceutical attributes affected by cocrystal formation and its relation to coformer properties.
Figure 4. Pharmaceutical attributes affected by cocrystal formation and its relation to coformer properties.
Pharmaceutics 15 01161 g004
Figure 5. Layered cocrystal structure of epalrestat and betaine. Reprinted (adapted) with permission from [21]. Copyright (2018) American Chemical Society.
Figure 5. Layered cocrystal structure of epalrestat and betaine. Reprinted (adapted) with permission from [21]. Copyright (2018) American Chemical Society.
Pharmaceutics 15 01161 g005
Figure 6. Intrinsic dissolution rate experiment results of epalrestat and its cocrystal. Reprinted (adapted) with permission from [21]. Copyright (2018) American Chemical Society.
Figure 6. Intrinsic dissolution rate experiment results of epalrestat and its cocrystal. Reprinted (adapted) with permission from [21]. Copyright (2018) American Chemical Society.
Pharmaceutics 15 01161 g006
Figure 7. Illustration of the influence of coformer solubility on the spring and parachute effect of cocrystals.
Figure 7. Illustration of the influence of coformer solubility on the spring and parachute effect of cocrystals.
Pharmaceutics 15 01161 g007
Figure 8. Role of coformers in stabilizing the drug molecule.
Figure 8. Role of coformers in stabilizing the drug molecule.
Pharmaceutics 15 01161 g008
Figure 9. Cumulative amount permeated for (A) Cocrystal, (B) Physical mixtures of 5-fluorouracil. Reprinted (adapted) with permission from [78]. Copyright (2016) American Chemical Society.
Figure 9. Cumulative amount permeated for (A) Cocrystal, (B) Physical mixtures of 5-fluorouracil. Reprinted (adapted) with permission from [78]. Copyright (2016) American Chemical Society.
Pharmaceutics 15 01161 g009
Figure 10. Illustration of enhanced permeability of cocrystals in comparison to pure drug.
Figure 10. Illustration of enhanced permeability of cocrystals in comparison to pure drug.
Pharmaceutics 15 01161 g010
Figure 11. Graphical representation of coformers in high demand in the last 20 years by (A) ScienceDirect (B) Web of Science, and (C) PubMed.
Figure 11. Graphical representation of coformers in high demand in the last 20 years by (A) ScienceDirect (B) Web of Science, and (C) PubMed.
Pharmaceutics 15 01161 g011
Figure 12. Literature trend of carboxylic acid-based coformers usage in preparation of cocrystals.
Figure 12. Literature trend of carboxylic acid-based coformers usage in preparation of cocrystals.
Pharmaceutics 15 01161 g012
Figure 13. Chemical structure of fumaric acid (A) 2d and (B) 3d.
Figure 13. Chemical structure of fumaric acid (A) 2d and (B) 3d.
Pharmaceutics 15 01161 g013
Figure 14. Hydrogen bond interaction and bond length between drug and coformer represented as (A) Promethazine hydrochloride and fumaric acid (#1853757), (B) Fluoxetine hydrochloride and fumaric acid (#254849), and (C) Meloxicam and fumaric acid (#819114) (from Mercury 2022.2.0 (Build 353591)). #—CCDC Identifier number.
Figure 14. Hydrogen bond interaction and bond length between drug and coformer represented as (A) Promethazine hydrochloride and fumaric acid (#1853757), (B) Fluoxetine hydrochloride and fumaric acid (#254849), and (C) Meloxicam and fumaric acid (#819114) (from Mercury 2022.2.0 (Build 353591)). #—CCDC Identifier number.
Pharmaceutics 15 01161 g014
Figure 15. Chemical structure of oxalic acid (A) 2d and (B) 3d.
Figure 15. Chemical structure of oxalic acid (A) 2d and (B) 3d.
Pharmaceutics 15 01161 g015
Figure 16. Hydrogen-bond interaction and bond length between drug and coformer represented as (A) Temozolomide and oxalic acid (#881197), (B) Promethazine hydrochloride and oxalic acid (#1853755), and (C) Caffeine and oxalic acid (#272620) (from Mercury 2022.2.0 (Build 353591)). #—CCDC Identifier number.
Figure 16. Hydrogen-bond interaction and bond length between drug and coformer represented as (A) Temozolomide and oxalic acid (#881197), (B) Promethazine hydrochloride and oxalic acid (#1853755), and (C) Caffeine and oxalic acid (#272620) (from Mercury 2022.2.0 (Build 353591)). #—CCDC Identifier number.
Pharmaceutics 15 01161 g016
Figure 17. (A) Chemical structure of succinic acid (A) 2d and (B) 3d [175].
Figure 17. (A) Chemical structure of succinic acid (A) 2d and (B) 3d [175].
Pharmaceutics 15 01161 g017
Figure 18. Hydrogen bond interaction and bond length between drug and coformer, represented as (A) Urea and Succinic acid (#795527), (B) Fluoxetine and Succinic acid (#254848), (C) Itraconazole and Succinic acid (#218567), and (D) Carbamazepine and Succinic acid (#671498) (from Mercury 2022.2.0 (Build 353591)). #—CCDC Identifier number.
Figure 18. Hydrogen bond interaction and bond length between drug and coformer, represented as (A) Urea and Succinic acid (#795527), (B) Fluoxetine and Succinic acid (#254848), (C) Itraconazole and Succinic acid (#218567), and (D) Carbamazepine and Succinic acid (#671498) (from Mercury 2022.2.0 (Build 353591)). #—CCDC Identifier number.
Pharmaceutics 15 01161 g018
Figure 19. Chemical structure of oxalic acid (A) 2d and (B) 3d.
Figure 19. Chemical structure of oxalic acid (A) 2d and (B) 3d.
Pharmaceutics 15 01161 g019
Figure 20. Hydrogen bond interaction and bond length between drug and coformer represented as (A) Paracetamol and citric acid (#803736), (B) Praziquantel and citric acid (#2094551), and (C) Nitrofurantoin and citric acid (#835423) (from Mercury 2022.2.0 (Build 353591)). #—CCDC Identifier number.
Figure 20. Hydrogen bond interaction and bond length between drug and coformer represented as (A) Paracetamol and citric acid (#803736), (B) Praziquantel and citric acid (#2094551), and (C) Nitrofurantoin and citric acid (#835423) (from Mercury 2022.2.0 (Build 353591)). #—CCDC Identifier number.
Pharmaceutics 15 01161 g020
Figure 21. An illustration of steps involved in cocrystal-based product development.
Figure 21. An illustration of steps involved in cocrystal-based product development.
Pharmaceutics 15 01161 g021
Table 1. Examples of studies reported on the effects of coformers on the solubility, dissolution rate, and bioavailability of drug molecules in cocrystals.
Table 1. Examples of studies reported on the effects of coformers on the solubility, dissolution rate, and bioavailability of drug molecules in cocrystals.
DrugCoformerEffect on Solubility (Comparison with Drug)Effect on Dissolution (Comparison with Drug)MechanismReference
EpalrestatBetaine2-fold increase3.5-fold increasePresence of layered structure with alternate drug and coformer[21]
GlibenclamideHippuric acid2.2-fold increase2.2-fold increaseHydrogen-bonding interactions between drug and coformers
Melting point of cocrystal Solubility of coformer
[48]
Nicotinic acid3-fold increase3-fold increase
Theophylline1.5-fold increase1.6-fold increase
Succinic acid3.5-fold increase3.7-fold increase
Diclofenacl-proline7.69-fold increase-Formation of layered structure[47]
ResveratrolPiperazine2.48-fold increase-Hydrogen bond network[49]
Methenamine1.46-fold increase
FelodipineGlutaric acid-2.41-fold increase at pH 1.2Lowering of melting point[50]
FurosemideIsonicotinamide5.6-fold increase (Apparent Solubility-Phase transformation)No changeCoformer solubility[51]
Eprosartan mesylateSalicylic acid6-fold increaseIncreased IDRCoformer solubility
Hydrogen-bonding interaction
[52]
p-aminobenzoic acid32-fold increase
Succinic acid61-fold increase
AcyclovirFumaric acid5.5-fold increaseIncreased dissolution rateCoformer solubility
Melting points of cocrystal
[53]
Malonic acid5.7-fold increaseIncreased dissolution rate
DL-tartaric acid5.3-fold increaseIncreased dissolution rate
HydrochlorothiazideNicotinic acid0.72 decrease-Formation of higher polarity cocrystals[54]
Nicotinamide1.27 increase
4-aminobenzoic acid0.21 decrease
Succinamide2.4-fold increase
Resorcinol1.3-fold increase
ItraconazoleAspartic acid3.2-fold increaseImproved dissolution rateHydrogen-bonding interaction[55]
Proline2.2-fold increase
Serine2.5-fold increase
Glycine2.3-fold increase
Succinic acid1.6-fold increase
PyrazinamideAdipic acid0.7-fold decrease0.6-fold decreaseHydrogen-bonding interaction[56]
Sebacic acid0.7-fold decrease0.4-fold decrease
trans-Aconitic acid1.6-fold increase1.8-fold increase
Citric acid1.2-fold increase1.4-fold increase
- Not mentioned.
Table 2. Examples of studies reported on mechanical properties of drug molecules in cocrystals.
Table 2. Examples of studies reported on mechanical properties of drug molecules in cocrystals.
DrugCoformerIssue of Drug/Objective of the StudyParameter ImprovedInferenceReference
CaffeineMethyl gallateSevere lamination and over compactionPowder compaction and tensile strengthPresence of slip planes resulted in improved compaction properties[59]
Paracetamol form II
  • Oxalic acid
  • Theophylline
  • Naphthalene
  • Phenazine
Poor tablet-forming abilityTensile strength, elastic constant and lattice energies, elastic compliance tensorLayered structure of cocrystals leads to superior tablet formation ability[58]
TheophyllineMethyl gallateTo examine the effects of cocrystallization on crystal mechanical propertiesElastic modulus, indentation values, crystal slip planes and Burger’s vectorTableting performance theophylline > co-crystal > methyl gallate[57]
Ibuprofen and FlurbiprofenNicotinamideTo demonstrate improvement of pharmaceutical properties over pure drug crystalPowder compaction analysis (tensile strength)Tabletability of cocrystal is apparently higher due to its higher bonding strength.[63]
ParacetamolTrimethyl glycinePoor tablet-forming abilityHardness and particle-size distributionImproved compression properties due to structural stability and changed crystal face[64]
Voriconazole
  • Fumaric acid
  • 4-hydroxybenzoic acid
  • 4-aminobenzoic acid
  • Oxalate salt
Too soft for tableting and compactingNanoindentation, elastic modulus and hardnessHardness improved in the order of oxalic acid salt > cocrystals > drug[65]
Lamotrigine
  • Cinnamic acid
  • Ferulic acid
  • Salicylic acid
  • Vanillic acid
Poor flow properties and cappingFlow (angle of repose) and compression propertiesAll cocrystals except with ferulic acid showed improved flow properties; capping exhibited by cocrystal with salicylic acid was weak than that of ferulic acid[66]
Paracetamol5-nitroisophthalic acidPoor mechanical propertiesTabletability – tensile strengthSuperior tabletability of cocrystal than the drug[67]
Sulfadimidine4-aminosalicylic acidTo investigate the cocrystal habit engineering effect on compaction propertiesDensity and Carr’s compressibility indexCrystal habit engineering of cocrystals leads to improved flow properties[68]
Adefovir DipivoxilStearic acidTo investigate the enhanced powder properties of cocrystalCompressibility (powder rheology analysis), Heckel analysisTabletability enhanced due to altered crystal habit by coformer[69]
Table 3. Examples of studies reporting on coformers’ effect on stabilization of drug molecules in cocrystals.
Table 3. Examples of studies reporting on coformers’ effect on stabilization of drug molecules in cocrystals.
DrugCoformer(s)Stability Issue of DrugParameter AssessedInferenceReference
Caffeine
  • Maleic acid
  • Oxalic acid
  • Glutaric acid
  • Malonic acid
Crystalline powder of anhydrous caffeine transforms to caffeine hydrate at high RHPhysical stability at storage conditions of 0, 43, 75, and 98% RH up to 7 weeks.No cocrystal hydrates have been found. Only oxalic acid-cocrystal exhibited physical stability till 7 weeks. The rest of cocrystals dissociated during storage. [71]
Nitrofurantoin4-hydroxybenzoic acidPhotosensitive and physicochemically unstablePhysical, chemical, and photostability at different conditions for 13 weeksImproved physicochemical and photostability compared to pure drug[73]
Adefovir dipivoxil
  • Saccharin
  • Nicotinamide
Degradation by hydrolysis and dimerization during storageChemical stability at 60 °C (60% RH) and 40 °C (75% RH) for a monthSaccharin cocrystal was stable for one month whereas nicotinamide was not stable[22]
Temozolomide
  • Salicylic acid
  • Oxalic acid
  • d,l-maleic acid
  • Succinic acid
  • d,l-tartaric acid
Spontaneous degradation during storage under normal conditions and is stable at pH < 5 but labile at pH > 7Chemical stability at 40 °C and 75% RH for 28 weeksInhibited the hydrolytic degradation of the drug as cocrystal by providing the acidic environment with organic acid coformers[74]
Tranilast
  • Urea
  • Nicotinamide
Photochemically unstableSubjected to 25 °C and 60% for 96 hPhotostability improved after the formation of cocrystal due to the increase in the distance between the drug molecules in the cocrystal[75]
Acyclovir
  • Fumaric acid
  • Malonic acid
  • Tartaric acid
Hydration of drug during storagePhysical stability at storage conditions of 0%, 43%, 75%, and 98% RH for 3 weeksCocrystal showed improved stability except with tartaric acid[53]
Etoricoxib
  • Suberic acid
  • Glutaric acid
  • Adipic acid
  • Caprolactam
Hemihydrate conversion of drug during manufacturing or upon exposure to moisture (30 min)Exposed to water (slurry) conditions for hydrationFormed stable cocrystals by replacing the water molecule in the crystal lattice[72]
EpalrestatBetainePhoto instabilitySubjected to 25 °C for 24 hImproved photostability of cocrystal due to decreased reaction cavity[21]
Isoniazid
  • Vanillic acid
  • Caffeic acid
Reaction of isoniazid and rifampicin in fixed-dose combinationPhysical stability of isoniazid in accelerated conditionsStronger hydrogen bond interaction and cyclic O-H···O synthon in the crystal structure stabilized the cocrystal.[76]
Flucytosine
  • Gallic acid
  • Glutaric acid
Susceptible to hydrationSubjected to 70–75% RH and 90–95% RH at ambient temperatureStable cocrystal may be due to strong acid–amide heterosynthon between drug and coformer[77]
Table 5. Reported coformers in literature used in the formation of cocrystals.
Table 5. Reported coformers in literature used in the formation of cocrystals.
API NameCoformer/API NameCocrystalsAPI:Coformer RatioReference
Temozolomide (TMZ)Nicotinamide (NCT)TMZ-NCT Cocrystal2:1[85]
Isonicotinamide (INA)TMZ-INA Cocrystal2:1
Pyrazinamide (PYZ)TMZ-PYZ Cocrystal1:1
Saccharin (SAC)TMZ-SAC Cocrystal2:1
Caffeine (CAF)TMZ-CAF Cocrystal1:1
Aripiprazole (ARI)Orcinol (ORC)ARI-ORC Cocrystal1:1[86]
Catechol (CAT)ARI-CAT Cocrystal1:1
Resorcinol (RES)ARI-RES Cocrystal1:1
Phloroglucinol (PHL)ARI-PHL Cocrystal1:1
Favipiravir (FAV)4-hydroxybenzoic acid (4HBA)FAV-4HBA Cocrystal1:1[87]
p-aminobenzoic acid (PABA)FAV-PABA Cocrystal1:1
Ferulic acid (FRA)FAV-FRA Cocrystal1:1
Gallic acid (GA)FAV-GA Cocrystal1:1
p-aminosalicylic acid (PAS)Pyrazine (PYZ)PAS-PYZ Cocrystal1:1[42]
Pyrimidine (PYM)PAS-PYM Cocrystal1:1
Pyridazine (PDZ)PAS-PDZ Cocrystal2:1
Phenazine (PHZ)PAS-PHZ Cocrystal1:2
4,4′-dipyridyl disulfide (DPDS)PAS-DPDS Cocrystal1:1
4-cyanopyridine (4-CYP)PAS-4-CYP Cocrystal (9)1:1
Curcumin (CUR)Salicylic acid (SAA)CUR-SAA Cocrystal1:2[88]
Hydroxyquinol (HXQ)CUR-HXQ Cocrystal1:1, 1:2
Resorcinol (RNL)CUR-RNL Cocrystal1:1
Pyrogallol (PYG)CUR-PYG Cocrystal1:1
4,4′-bipyridine N, N′-dioxide (4,4 BPDO)CUR-4,4 BPDO Cocrystal
Salicylic acid (SA)Benzamide (BZ)SA-BZ Cocrystal1:1, 1:2
Isonicotinamide (INA)SA-INA Cocrystal1:1, 2:1
Carbamazepine (CMP)4-aminobenzoic acid (4, ABA)CMP-4 ABA Cocrystal1:1, 2:1, 4:1
Nicotinamide (NCT)r-mandelic acid (r-MDLA)NCT-r-MDLA Cocrystal1:2, 1:1, 4:1
Urea (UA)Succinic acid (SA)UA-SA Cocrystal1:1, 2:1
Indomethacin (IMC)Saccharin (SAC)IMC-SAC Cocrystal
CL-20Pyrazine (PYZ)CL-20-PYZ Cocrystal [23]
Theobromine (TBR)Trimesic acid (TMSA)TBR-TMSA Cocrystal1:1[89]
Theophylline (TPH)Trimesic acid (TMSA)TPH-TMSA Cocrystal1:1
Caffeine (CAF)Trimesic acid (TMSA)CAF-TMSA Cocrystal1:2
Theobromine (TBR)Hemimellitic acid (HMLA)TBR-HMLA Cocrystal1:1
Theophylline (TPH)Hemimellitic acid (HMLA)TPH-HMLA Cocrystal1:1
Caffeine (CAF)Hemimellitic acid (HMLA)CAF-HMLA Cocrystal10:1
5-Fluorouracil (5-FU)Succinic acid (SA)5-FU-SA Cocrystal1:1[90]
Phenazine (PHZ)5-FU-PHZ Cocrystal2:1
Acridine (ACD)5-FU-ACD Cocrystal2:1
Benzoic acid (BA)5-FU-BA Cocrystal1:1
Malic acid (MA)5-FU-MA Cocrystal1:1
Cinnamic acid (CA)5-FU-CA Cocrystal1:1
4,4-bispyridylethene (4,4 BPYE)5-FU-4,4 BPYE Cocrystal4:1
p-aminopyridine (p-APY)5-FU-p-APY Cocrystal
Rivaroxaban (RVB)Malonic Acid (MA) & Oxalic Acid (OA)RVB-MA Cocrystals and RVB-OA Cocrystals1:1, 2:1[91]
p-hydroxybenzoic acid (pHBA) RVB-pHBA Cocrystals1:1[92]
Isonicotinamide (INTA)RVB-INTA Cocrystal1:1
Nicotinamide (NTA)RVB-NTA Cocrystal1:1
2,4 dihydroxybenzoic acid (2,4 DHBA) RVB-2,4 DHBA Cocrystals1:1
Succinic acid (SA)RVB-SA Cocrystals1:1
Malonic Acid (MA)RVB-MA Cocrystals2:1[93]
Malonic Acid (MA)RVB-MA Cocrystals2:1[94]
Lamotrigine (LTG)Phthalimide (PTA)LTG-PTA Cocrystals1:1[95]
Succinic acid (SA)LTG-SA Cocrystals
Pyromellitic diimide (PDA)LTG-PDA Cocrystals1:1[96]
Caffeine (CAF)LTG-CAF Cocrystals2:1
Isophthaldehyde (IPA)LTG-IPA Cocrystals1:1
Glutarimide (GTA)LTG-GTA Cocrystals1:1[97]
Phenobarbital (PBT) (Multi drug cocrystals)LTG-PBT Cocrystals1:3, 3:1[98]
2,2′-bipyridine (2,2 BPYD)LTG-2,2 BPYD Cocrystals1:1.5[99]
4,4′-bipyridine (4,4 BPYD)LTG-4,4 BPYD Cocrystals2:1
Nicotinamide monohydrate (NTAM)LTG-NTAM Cocrystal1:1:1[100]
Acetamide (ACT)LTG-ACT Cocrystal1:1
Acetic acid (ATA)LTG-ATA Cocrystal1:3
4-hydroxy-benzoic acid (4 HBA)LTG-4 HBA Cocrystal1:1
Saccharin (SAC)LTG-SAC Cocrystal1:1
Etodolac (ETD)4-amino benzoic acid (4 ABA)ETD-4 ABA Cocrystal1:1[101]
Glutaric acid (CA)ETD-GA Cocrystal1:2[102]
Table 6. List of commercialized cocrystals.
Table 6. List of commercialized cocrystals.
Drug NameApprovalComponentsDosage FormIndicationManufacturerRef. No.
Depakote®U.S. FDA 1983Valproic acid + Valproate sodium Tablet, CapsuleEpilepsyAbbott Laboratories, Illinois, United States[15,16,103,104,105,106]
Entresto®U.S. FDA 2015Sacubitril sodium + Valsartan sodium TabletHeart failureNovartis, Basel, Switzerland[15,16,107]
Suglat®Japan 2014Ipragliflozin + L-prolineTabletDiabetesKotobuki Pharmaceuticals, Nishina, Shizuoka, Japan and Astellas Pharma, Tokyo, Japan [15,16,105,108]
Steglatro®U.S. FDA 2017Ertugliflozin + L-pyroglutamic acidTabletDiabetesPfizer, New York, United States[15,105,109,110]
Lexapro®U.S. FDA 2002Escitalopram oxalate + Oxalic acidTabletAnxiety and depressionAllergan Inc., Dublin, Ireland[15,16,111,112]
ESIX-10®U.S. FDA 2009Escitalopram oxalate + Oxalic acidTabletAnxiety and depression Sag Health Science Pvt Ltd., New Delhi, India
Beta-chlor®U.S. FDA 1963Chloral hydrate + BetaineTabletSedationMead Johnson, Illinois, United States[15,105]
Cafcit®U.S. FDA 1999 Caffeine + Citric acid InjectionInfantile apnoeaHikma Pharmaceuticals Plc, London, United Kingdom[16,105,113]
Zafatek®Japan
2015
Trelagliptin + Succinic acidTabletDiabetesTakeda Pharmaceutical Company Limited, Tokyo, Japan [16,114]
Lamivudine/zidovudine Teva ®EMA
2011
Lamivudine + ZidovudineTabletHIV infectionTeva Pharma B.V., Tel Aviv-Yafo, Israel[16,115,116,117]
Abilify ®U.S. FDA 2002Aripiprazole + Fumaric acidTabletSchizophreniaOtsuka Pharmaceuticals, Tokyo, Japan[128,129,135]
Odomzo®U.S. FDA 2015Sonidegib + Phosphoric acidCapsuleBasal Cell CarcinomaSun Pharma Global, Mumbai, India.[118,119,120]
Mayzent®U.S. FDA 2019Siponimod + Fumaric acidTabletMultiple SclerosisNovartis, Basel, Switzerland[121,122,123]
Seglentis®U.S. FDA 2021Celecoxib + Tramadol TabletAcute PainKowa Pharmaceuticals, Alabama, United States[122,124]
DimenhydrinateU.S. FDA 1982 (ANDA)Diphenhydramine and 8-chlorotheophyllineTabletMotion sicknessWatson Laboratories Inc., New Jersey, United States[17,125,126]
Ibrutinib fumaric acid cocrystalsTentative approvalIbrutinib + Fumaric acidNACancerTeva Pharmaceutical Industries Ltd., Tel Aviv-Yafo, Israel[15,127]
E-58425 (Clinical Trial Phase 3)Approval PendingCelecoxib and racemic tramadol hydrochlorideNAManagement of acute painPatented by Laboratorios Del., La Paz, Bolivia Development done by Enantia and Esteve, R&D, Spain[128,129,130,131,132]
TAK-020 (Clinical Trial Phase 1)Approval PendingTAK-020 and Gentisic acidNARheumatoid arthritisTakeda Pharmaceuticals, Tokyo, Japan[128,129,133,134]
Table 7. Physical and chemical properties of carboxylic acid-based coformers.
Table 7. Physical and chemical properties of carboxylic acid-based coformers.
CoformerFumaric Acid [136,137,138,139]Oxalic Acid [91,140,141]Succinic acid [142]Citric Acid [143,144]
Physical state Colorless crystalline solid Colorless crystalline solid Colorless, odorless white crystals Colorless crystalline solid
Melting point287 °C189.5 °C185–187 °C153 °C
Solubility in solventsSoluble in ethanol, concentrated sulfuric acid.
Slightly soluble in ethyl ether, acetone.
Insoluble in chloroform and benzene.
Very soluble in ethanol.
Slightly soluble in ether.
Insoluble in benzene, chloroform, and petroleum ether.
Slightly soluble in ethanol, ether, acetone, glycerin.
Not soluble in benzene, carbon sulfide, carbon tetrachloride.
Freely soluble in ethanol.
Insoluble in benzene, chloroform, carbon tetrachloride, toluene, and carbon disulfide.
Solubility in water7 g/L (25 °C)220 mg/mL (25 °C)Soluble (71 mg/mL)592 mg/mL (20 °C)
Molar mass 116.07 90.03 118.09192.1
Density1.64 g/cm31.9 g/cm31.56 g/cm31.66 g/cm3
pKa 3.03 1.2 4.242.79
No. of hydrogen bond donors2224
No. of hydrogen bond acceptors4447
StabilityStable under ambient conditionsStable under ambient conditionsStable under ambient conditions Moisture-sensitive
Table 8. Impact of fumaric acid as coformer on drugs.
Table 8. Impact of fumaric acid as coformer on drugs.
CocrystalMethod of PreparationImpact on SolubilityImpact on Dissolution RateImpact on BioavailabilityImpact on StabilityReferences
Berberine–Fumaric acid
(2:1)
Slurry method~9.5-fold at 15 min #~3.75-fold────Reduced hygroscopicity[7]
Promethazine hydrochloride–Fumaric acid
(2:1)
MechanochemistryImproved────────Improved[146]
Slow solvent evaporation
Gabapentin-lactam–Fumaric acid (1:1)Reaction crystallization methodImproved────────────[147,158]
Fluoxetine HCl–Fumaric acid (2:1)Cooling crystallizationIncreased
~2-fold
No improvement────────[151]
Enoxacin–Fumaric acid
(1:2) @
Slow solvent evaporation9.8-fold8.9-fold────────[150]
Meloxicam–Fumaric acid
(1:1)
Liquid-assisted grinding33–84% improvement────────────[159]
Sildenafil–Fumaric acid
(1:2 and 1:3)
Slow solvent evaporationIncreased 5-fold────────────[156]
PEC–Fumaric acid
(1:1)
Anti-solvent additionIncreased 4-foldImprovedImproved────[160]
Glipizide–Fumaric acid
(1:1)
Liquid-assisted grindingIncreased ~2.3-foldIncreased
~2-fold
Improved────[161]
Acyclovir–Fumaric acid
(1:1)
Slow solvent evaporation & liquid-assisted grindingIncreased
~5.5-fold
Increased
~2-fold
────Improved[53]
Ketoconazole–Fumaric acid
(1:1, 1:2, and 1:3)
Slow solvent evaporationIncreased
~1.6-fold
Improved ~1.65-fold────Improved[157]
Acyclovir–Fumaric acid
(1:1)
Dry grinding or co-grindingLess effectIncreased
~2.2-fold
────────[162]
Ethenzamide–Fumaric acid
(2:1)
Slow solvent evaporationIncreased 3.84-foldIncreased 1.71-fold────────[163]
Fluconazole–Fumaric acid
(1:1)
Slow solvent evaporationIncreased ~2.5-fold #Improved────Improved[164]
Efavirenz–Fumaric acid
(1:1)
Neat grindingIncreased
~26-fold
Increased
~2-fold
────────[165]
#—Equilibrium solubility; @—2.3-fold improvement in permeability was also reported.
Table 9. Impact of oxalic acid as a coformer on drugs.
Table 9. Impact of oxalic acid as a coformer on drugs.
CocrystalMethod of PreparationImpact on SolubilityImpact on Dissolution rateImpact on BioavailabilityImpact on StabilityReferences
Promethazine HCl–Oxalic acid (2:1)Slow solvent evaporationImproved────────Improved[146]
Telmisartan–Oxalic acid────Increased 7-foldIncreased
~2.4-fold
────────[122,169]
Rebamipide–Oxalic acid (1:1)Liquid-assisted grindingIncreased 7.29-foldIncreased
7.19-fold
Increased
1.6-fold
────
Rivaroxaban–Oxalic acid (1:1)Anti-solvent additionImprovedIncreased
~1.6-fold
Increased
~2.12-fold
────[91]
Apixaban–Oxalic acid
(4:3)
────Increased approx. 2-fold────Enhanced 2.7-fold────[166]
Temozolomide–Oxalic acid (2:1)────────────────Improved[167]
Xanthotoxin–Oxalic acid (2:1)Liquid assisted grinding & slow solvent evaporationIncreased 1.6-foldIncreased
~1.1-fold
────Improved[168]
Telmisartan–Oxalic acid (1:1)Solvent-drop grinding & solvent evaporation methodIncreased 11.7-foldIncreased
~7.2-fold
────────[170]
Caffeine–Oxalic acid
(2:1)
Solvent precipitation and ultrasound-assisted solution cocrystallization────────────Improved[171]
Glibenclamide–Oxalic acid (1:2)Thermal methodIncreased
~2.7-fold
Increased
~1.7-fold
────────[172]
Table 10. Impact of succinic acid as a coformer on drugs.
Table 10. Impact of succinic acid as a coformer on drugs.
API-CoformerMethod of PreparationImpact on SolubilityImpact on Dissolution RateImpact on BioavailabilityImpact on StabilityReferences
Itraconazole–Succinic acidLiquid anti-solvent────F1—Achieved 50% release in 2 h────Improved[178]
Gas anti-solventF2—Achieved 92% release in 2 h
Piperine–Succinic acid Wet-milling Increased 12.70-foldAchieved 53.281% release in 1 h────Improved[179]
Carbamazepine–Succinic acid (2:1)Slurry crystallizationImprovedF1—Achieved 82% release in 1 h
F2—Achieved 95% release in 1 h
F3—Achieved 95% release in 1 h
ImprovedImproved[2]
Isoniazid–Succinic acid
(2:1)
Slow solvent evaporation────Improved────────[180]
Imidazopyridazine- Succinic acid (1:1)Neat grinding methodImprovedImproved────────[181]
Brexpiprazole–Succinic acid (1:1)Solvent-drop grinding methodIncreased 1.59-fold────────────[182]
Aripiprazole–Succinic acidHot melt extrusion (HME)────Improved────────[183]
Ketoconazole–Succinic acid (1:1)Reaction crystallization method────Decreased with the low pH of coformer────────[184]
Abiraterone acetate–Succinic acid (2:1)Solvent evaporation────Increased 4.7-fold────Improved[185]
Acyclovir–Succinic acid
(1:1)
Grinding method Dissolution efficiency– 54.23% (grinding time 15 min)────────[186]
Slurry crystallization Dissolution efficiency – 74.36% (solvent concentration 12 mL/g)────────
Fluoxetine HCl–Succinic acid (2:1)Slow solvent evaporationIncreased
~1.5-fold
Increased
3-fold
────────[151]
F1—Formulation 1, F2—Formulation 2, F3—Formulation 3.
Table 11. Impact of citric acid as a coformer on drugs.
Table 11. Impact of citric acid as a coformer on drugs.
CocrystalsMethod of PreparationImpact on SolubilityImpact on Dissolution rateImpact on BioavailabilityImpact on StabilityReferences
Rebamipide–Citric acid
(1:1)
Liquid-assisted grindingIncreased 12.58-foldIncreased ~13.2-foldIncreased
2.5-fold
────[169]
Metformin hydrochloride–Citric acid (1:1)Solution crystallization, neat grinding, and liquid-assisted grindingIncreased 1–4-fold────Improved────[143]
Berberine chloride- Citric acid (1:1)Liquid-assisted grindingImproved────────Improved[6]
Theophylline–Citric acid
(1:1)
Neat co-grinding────────────Improved[189]
Dapagliflozin propanediol monohydrate–Citric acid
(1:1)
Solution crystallization methodImproved Increased
~1-fold
────Improved[190]
Pyrazinamide–Citric acid
(1:1)
Slow solvent evaporationIncreased
~1.1-fold
Increased
~1.4-fold
────────[56]
Ethyl p-methoxycinnamate–Citric acid (1:1, 1:2, 1:3)Liquid-assisted grindingIncreased
1.4-fold
────────────[192]
Simvastatin–Citric acid
(1:1)
Liquid-assisted grinding, slow solvent evaporationIncreased
1.4–3-fold
────ImprovedImproved[193]
Nefiracetam–Citric acid
(2:1)
Slow solvent evaporationIncreased
~1.4-fold
Improved ────────[195]
Ritonavir–Citric acid
(1:2)
Dry grinding methodImprovedImproved────────[196]
Nitrofurantoin–Citric acid
(1:1)
Liquid-assisted grindingImproved────────Improved[197]
Praziquantel–Citric acid
(1:1)
Liquid-assisted grindingIncreased
~2- 4-fold
Improved────────[198]
Table 12. Role of coformers on the physicochemical properties of drug.
Table 12. Role of coformers on the physicochemical properties of drug.
S. No.ParameterFumaric AcidOxalic AcidSuccinic AcidCitric Acid
1.SolubilityImprovedImprovedImprovedImproved
2.Dissolution rateImprovedImprovedImprovedImproved
3.PermeabilityImprovedNo ImpactNo ImpactNo Impact
4.BioavailabilityImprovedImprovedImprovedImproved
5.StabilityImprovedImprovedImprovedImproved
6.No. of commercialized cocrystalsTwoTwoOneOne
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Singh, M.; Barua, H.; Jyothi, V.G.S.S.; Dhondale, M.R.; Nambiar, A.G.; Agrawal, A.K.; Kumar, P.; Shastri, N.R.; Kumar, D. Cocrystals by Design: A Rational Coformer Selection Approach for Tackling the API Problems. Pharmaceutics 2023, 15, 1161. https://doi.org/10.3390/pharmaceutics15041161

AMA Style

Singh M, Barua H, Jyothi VGSS, Dhondale MR, Nambiar AG, Agrawal AK, Kumar P, Shastri NR, Kumar D. Cocrystals by Design: A Rational Coformer Selection Approach for Tackling the API Problems. Pharmaceutics. 2023; 15(4):1161. https://doi.org/10.3390/pharmaceutics15041161

Chicago/Turabian Style

Singh, Maan, Harsh Barua, Vaskuri G. S. Sainaga Jyothi, Madhukiran R. Dhondale, Amritha G. Nambiar, Ashish K. Agrawal, Pradeep Kumar, Nalini R. Shastri, and Dinesh Kumar. 2023. "Cocrystals by Design: A Rational Coformer Selection Approach for Tackling the API Problems" Pharmaceutics 15, no. 4: 1161. https://doi.org/10.3390/pharmaceutics15041161

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