Next Article in Journal
Atomic-Site-Specific Analysis on Out-of-Plane Elasticity of Convexly Curved Graphene and Its Relationship to s p 2 to s p 3 Re-Hybridization
Previous Article in Journal
Incorporation of Hexanuclear Mn(II,III) Carboxylate Clusters with a {Mn6O2} Core in Polymeric Structures
Previous Article in Special Issue
Febuxostat-Minoxidil Salt Solvates: Crystal Structures, Characterization, Interconversion and Solubility Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Drug‑Drug and Drug‑Nutraceutical Cocrystal/Salt as Alternative Medicine for Combination Therapy: A Crystal Engineering Approach

1
Department of Chemistry, Gauhati University, Guwahati-781014, Assam, India
2
Department of Chemical Sciences, Tezpur University, Tezpur-784028, Assam, India.
*
Authors to whom correspondence should be addressed.
Crystals 2018, 8(2), 101; https://doi.org/10.3390/cryst8020101
Submission received: 3 December 2017 / Revised: 1 February 2018 / Accepted: 12 February 2018 / Published: 18 February 2018
(This article belongs to the Special Issue Novel Pharmaceutical Cocrystals and Their Applications)

Abstract

:
The pre-formulation of pharmaceutical cocrystals and salts is a concept of crystal engineering that has emerged as a promising technique for drug development in pharmaceutical industry. Recent introduction of pharmaceutical cocrystals in regulatory guidelines of US Food and Drug Administration (FDA) made them one of the potential alternatives when salt preparation is not feasible. Apart from generally regarded as safe (GRAS) coformers, drug-drug and drug-nutraceutical cocrystals are recent additions to pharmaceutical cocrystal family that have additional health benefits. Indeed, preparation of salt forms is a routine practice to deal with inadequacies associated with the active pharmaceutical ingredient (API) and happens to be a potentially reliable method. Amongst them, drug-drug and drug-nutraceutical cocrystals have drawn significant importance in the recent past as they reduce drug load and cost effects during multiple disease diagnosis. However, one has to be prudent in the selection of drug molecules, the presence of complementary hydrogen bond synthon, disease management during multiple disease therapy, etc. that play important roles in their preparation. That is the reason why drug–drug cocrystals are scarce in the literature compared to pharmaceutical cocrystals containing GRAS coformers and salt forms. Herein, we discuss case studies preferably the reported drug-drug, drug-nutraceutical cocrystals, and a few salts with an emphasis on their role in physicochemical property modulation.

1. Introduction

The breakthrough of a new drug is time consuming and staggeringly difficult task. The discovery of new drugs is gradually decreasing because of stiffened regulations and requirements imposed on drug discovery and formulation [1,2]. The subject crystal engineering has brought us an opportunity to overcome these issues by translating the idea of new pre-formulations of existing drugs based on intermolecular interactions. Novel formulation with improved pharmacological properties has a vital necessity in this context because most drug molecules have more or less inadequacies relating to their pharmacokinetic and physiological properties such as solubility, dissolution, hygroscopicity, membrane permeation, chemical stability, etc. [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23]. They are imperative factors that essentially define the overall efficacy of the active pharmaceutical ingredient (API). One survey says that nearly 80% drugs are in solid formulations and absorbed via passive diffusion from the gastrointestinal tract [24]. Among them, over 40% of drugs that appear on the World Health Organization (WHO) Essential Drug List are known to be insoluble or poorly water-soluble, based on the Biopharmaceutics Classification System (BCS) [3,25]. Preparation of salt formulation made a drug more bio-available; however, more than 50% of drug molecules lack an ionisable group which can make salt formation likely impossible. From these facts, it is essential to develop a new solid state formulation with trusted properties. The brilliancy of the subject of crystal engineering is that it provides a wide-range of possibilities to engineer and/or design various crystalline solid state formulations of a drug that can carry the tailored biopharmaceutical properties—for example polymorphs, pseudopolymorphs, solvates, hydrates, cocrystal, etc. They have been demonstrated as alternatives and can easily minimize major obstacles in successfully commercializing the compounds. Among them, cocrystal synthesis is relatively modern technology to modulate drug biopharmaceutical properties, in particular to treat drug resistance issues [26,27,28,29]. Fortuitously, the importance of cocrystal preparation has recently been realized in the pharmaceutical industry and many of them have started practicing in order to modify an existing API for coveted attributes. This scope continues to develop a new formulation to enhance and/or control bioavailability and essentially the efficacy of a drug. A pharmaceutical cocrystal generally incorporates API with a pharmaceutically-acceptable molecule (cocrystal former or CCF) in their crystal lattice without any structural modification of the parent drug (Figure 1) [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,30,31,32,33,34,35]. Preferably, the coformers are chosen from the GRAS (generally regarded as safe) list prepared by USA Food and Drug Administration (FDA) which comprises over 3000 substances [36,37,38,39]. Food additives, preservatives, excipients, vitamins, minerals, amino acids, bio-molecules, and their APIs can be selected as CCF. They are commonly prepared by a variety of techniques such as melt crystallization, grinding, and recrystallization from solvents etc. Thus, pharmaceutical cocrystal stands here as an alternate and interesting pathway to improve efficacy of a drug.
Crystalline products are strongly preferred in the development of an API due to their relative ease in terms of purity, stability, and an established technology for product improvement synergistically by changing intermolecular interaction behaviors, such as drug-drug, drug-coformer and solute–solvent interactions. It is needless to mention that coformers and/or excipients play crucial roles in changing the lipophilic behaviors resulting modified physicochemical and mechanical properties. Other very important developments are to address drug resistance issues and combination therapy using cocrystals. Most bacteria, viruses, and other microbes multiply rapidly and slowly adapt resistance to the drugs designed to kill them, making the product less effective. Essentially, they develop a resistance power to that drug over time. Thus strategy to develop multicomponent drugs, especially when all components are APIs, would offer a combination drug that shows better effect than the sum of their individual effects [27,28,29,30,31,32]. A typical CCF when replaced by another drug or a nutraceutical, the resultant cocrystal is known as drug–drug and drug–nutraceutical cocrystal and they are now comprehended as alternate drugs for enhancing well-being. Down the line, we have discussed several examples from the context of pharmaceutical cocrystals that led to significant improvement in drug properties. Entresto is one of the classic examples to receive approval as a new active substance. It is a fixed-dose combination product presented as film-coated tablets containing and valsartan as active substances as a trisodium hemipentahydrate co-crystal. Generally, three strengths are proposed containing 24.3 mg sacubitril and 25.7 mg valsartan (low dose), 48.6 mg sacubitril and 51.4 mg valsartan (medium dose), and 97.2 mg sacubitril and 102.8 mg valsartan (high dose) for the treatment of chronic heart failure [40]. In 2009, Lexapro—which is an approved cocrystal/salt of Escitalopram and oxalic acid—was mainly used for the treatment of major depressive and anxiety disorders [41]. Though most pharmaceutical cocrystals remain in 1:1 equivalency, the stoichiometric ratio in product cocrystals is an important parameter to address. Our group studied synthesis of variable stoichiometry drug cocrystals that result with different solubilities and cell membrane permeations behaviour in vitro under physiological pH conditions [7]. In a few other systems, we further examined the essence of molecular symmetry and conformational behavior of drug in presence of supramolecular entities in altering pharmacokinetic properties [5,8,42]. From the plethora of publications in the area of pharmaceutical cocrystals, it has been noticed that most pharmaceutical cocrystals were formed between organic acids or bases and the API [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,30,31,32,33,34,35,43,44,45,46]. So there seems to remain a scope to explore cocrystal engineering approach using inorganic acid as a cocrystal former.
From thermodynamic point of view salts, cocrystals or amorphous forms are high energy forms [12]. They are widely used to generate supersaturated aqueous solutions and improve drug bioavailability. Generation and maintenance of the metastable supersaturated state (makes the drug more bioavailable) as a strategy to improve intestinal absorption of poorly water soluble drugs, there are two essential steps need to be considered and they are termed as ‘spring and parachute approach’ [12]. A thermodynamically unstable, supersaturated solution of a drug can only be generated starting from a high energy form of a drug which is known as the “spring” (Figure 2). A combination of excipients like co-solvent, lipid, or polymer based formulations; coformers etc. can deliver the drug in solution as high energy solid forms that can easily provide an accelerated dissolution or a higher apparent solubility. This phenomenon is known as the ‘parachute effect’. The apparent solubility is the apparent equilibrium between drug in solution and a solid whose structure is in the high energy state. A high energy form of the drug (the spring) provides the driving force to solubilize the drug at a concentration greater than its equilibrium solubility level and a similar effect resulted by the combination of excipients (the parachute) by inhibiting or retarding precipitation. Maintaining a supersaturated state for a time period sufficient for absorption may require a temporary inhibition of precipitation through the use of pharmaceutical excipients or by other components that interfere with nucleation and crystal growth. This are the ‘parachutes’ or precipitation inhibitors.
The strategy involves drug-coformer combinations that have the potential of forming energetically and structurally robust interactions which are highlighted as the use of hydrogen bonding and crystal engineering strategies to improve drug profile. In the context of drug–drug and drug–nutraceutical cocrystals, literature reveals only a few cocrystal case studies with pharmacokinetic details. However, R&D departments of many pharmaceutical companies have started putting effort into cocrystal synthesis as an alternative formulation. Only a very limited number of pharmaceutical cocrystals have been approved by the FDA as drug products to date [47,48,49,50,51,52,53,54]. Of late, on 7 July 2015, Novartis gained approval for Entresto™—a cocrystal comprised of sacubitril, valsartan, and solvate water—to treat chronic heart failure [41,55]. Cafcit, which is a drug product consisting anhydrous caffeine, citric acid monohydrate, and sodium citrate dehydrate is another cocrystal formulation already available in the market [56,57,58]. Sodium valproate is another marketed product with a drug-drug co-crystal [59]. Dimenhydrinate is another widely used drug for the prevention of motion sickness, including nausea and vomiting which is a cocrystal of two drugs—namely, diphenhydramine and 8-chlorotheophylline. The antiemetic effect of the drug for prevention of sickness is due to the presence of diphenhydramine, but associated with the side effect of drowsiness. This cocrystal of dimenhydrinate overcomes side effects as it contains an additional stimulant, 8-chlorotheophylline [60]. The recent approval of a pharmaceutical cocrystals has attracted additional interest from the pharmaceutical industry as materials for drug formulation. Thus novelty, non-obviousness, and usefulness can make cocrystals patentable from the perspective of intellectual property as well as drug formulation. A novel cocrystal formulation can emerge as viable alternative to the traditional drugs in near future. The patent portfolios of cocrystals as drug substance not drug product infer enhancement of existing commercial value of the product. The FDA perhaps issued revised guidance to technically distinguish between salts and cocrystals based on the ionization of the components. Essentially, it abandons the position that cocrystals are drug product intermediates. Cocrystals are therefore adapted as a special case of solvates, in which the second component is nonvolatile and refers to a conformer, rather than as an excipient.
The presence of many functional groups in APIs, which may form robust supramolecular synthons defined by Desiraju [61]. Later, Zaworotko sub-classified supramolecular synthons into homosynthons and heterosynthons based on the interacting functional groups [19]. Commonly observed acid…acid, acid…pyridine, acid…amide, amide…amide, amide…pyridine N-oxide, O−H···O, O−H···N, N−H···O, and N−H···N, etc. supramolecular synthons offer a great opportunity to design pharmaceutical cocrystals. Moreover, the role of molecular arrangement in the crystal lattice which is controlled by intermolecular interactions not only by hydrogen bonds—but also halogen bonds, π–π interactions [62], van der Waals interactions, metal-coordination bonding, etc.—on modifying properties of a solid crystalline substance also been studied.
As cited before salts and cocrystals can distinguish based on the ionization of the components. It is indeed essential to understand the classification of crystalline solid forms for making in order to characterize them and draw appropriate comparisons as both salts and cocrystals fall into multicomponent crystals. Cocrystals are defined as multicomponent crystal structure in which two or more compounds coexist through hydrogen bonds or non-covalent interactions and they are in neutral form. However, salt can be defined when the proton is transferred from acid to base and exists in the ionic state. Therefore, it is apparent that there exists a continuum linking cocrystals and salts based on the extent of proton transfer between acidic and basic components. When the proton resides on the base, then proton transfer has occurred and the crystalline acid–base complex is a salt and when the proton remains on the acid, then it is a cocrystal. The tendency of transferring a proton from the acidic end to the base is directly related to the pKa difference, which is defined as the negative logarithm of the dissociation constant [63,64,65,66,67,68,69,70,71]. Thus the ΔpKa [ΔpKa = pKa(base) − pKa(acid)] will decide the location of the transferable proton within the continuum. It has been observed that when the ΔpKa is more than 3.75, maximum probability of the proton will reside on the base and the resultant material is a salt, which is an essential criterion while selecting the appropriate counter ions to the preparation of salts of API in order to improve its properties like solubility. Similarly, a neutral cocrystal is expected when the pKa difference is less than 0. The proton will have an intermediate H bond character when the transition range falls between 0 and 3.75. This intermediate situation may lead to salt as well as cocrystal. A linear relationship between ∆pKa and the probability of proton transfer between acid-base pairs was also derived by Aurora from a CSD study of nearly 6500 crystalline complexes. The pKa rule was validated when ∆pKa is between −1 and +4 (Scheme 1b) [67]. The pKa rule of thumb to predict the proton location in a continuum is depicted in Scheme 1a. The other scale of pKHB which measures the free energy of hydrogen bonded complex (1.364 pKHB = −ΔGHB in kcal mol−1) also determines the hydrogen bond pairing that could be an appropriate scale compared to pKa values [72]. The pKHB values are sensitive to factors—like inductive effect, resonance effects, steric hindrance, lone-pair repulsion, intramolecular H-bonding, etc.—that can alter intermolecular interaction pattern.

2. Case Studies

2.1. Drug-Drug Cocrystal

2.1.1. Aspirin Cocrystals

Žegarac et al [73] prepared a 1:1 sildenafil–acetylsalicylic acid (more commonly known as aspirin) cocrystal and a 1:1:1 cocrystal salt involving sildenafil cation, acetylsalicylate and salicylic acid. Sildenafil is an active component of the blockbuster drug Viagra® used for the treatment of male erectile dysfunction. Solution crystallization resulted the 1:1 cocrystal along with accidental discovery of the 1:1:1 cocrystal salt. The 1:1 acetylsalicylic acid based cocrystal exhibits 75% improved dissolution rate compared to sildenafil citrate (the active ingredient of Viagra®) (Figure 3).
Meloxicam is a non-steroidal anti-inflammatory drug with low aqueous solubility and high permeability. In order to improve its aqueous solubility, Zaworotko and coworkers prepared a 1:1 cocrystal of meloxicam and aspirin [74]. The crystal structure contains a R 2 2 (8) ring motif connecting aspirin and meloxicam molecule as shown in Figure 4. The resulting cocrystal exhibits superior kinetic solubility and potentiality to significantly decrease the time required to reach the human therapeutic concentration compared with the parent drug, meloxicam.

2.1.2. Anti-Tuberculosis Drug Cocrystal

Desiraju and coworker reported [75] two drug-drug cocrystals of anti-tuberculosis drugs isoniazid (INH), pyrazinamide (Pyz), and 4-aminosalicylic acid (PAS). They were successful in preparing two drug-drug cocrystals—namely, 1:1 molecular complex of INH and PAS; and a monohydrate of 1:1 complex of Pyz and PAS—using liquid assisted grinding as well as solution crystallization. They observed that both the cocrystal system contains the most stable COOH···Npyridine hydrogen bond synthon. In case of INH-PAS cocrystal there are two symmetry independent COOH···Npyridine hydrogen bond present; one with H atom located on the carboxylic group indicative of pure cocrystal, whereas the second one shows a partial proton transfer. 3D crystal packing of INH-PAS cocrystal is shown in Figure 5. The two symmetry independent COOH···Npyridine hydrogen bond synthon present in INH-PAS cocrystal is shown in Figure 6. On the other hand, crystallizing Pyz with PAS in 1:1 ratio in methanol resulted a monohydrate of Pyz-PAS cocrystal. Various hydrogen bond synthon present in the cocrystal hydrate is shown in Figure 7. The water molecule actively takes part in the hydrogen bond connecting the Pyz and PAS molecules in a tetrahedral environment and acts simultaneously as two hydrogen bond donors and two hydrogen bond acceptors. Due to drug resistant TB, multidrug therapy is practiced during treatment of TB. Therefore, cocrystal of PAS with INH and Pyz may be a useful formulation for the treatment of tuberculosis.
Liu et al. [76] recently designed a drug-bridged-drug ternary cocrystallization strategy to prepare a combination drug utilizing a GRAS coformer fumaric acid to connect a Pyz and INH molecule as shown in Figure 8. This is the first ternary dual-drug cocrystal reported in the literature. They highlighted that the ternary cocrystal drug exhibited optimized formulation capacity and in vitro/vivo synergistic effects.

2.1.3. Bicalutamide Cocrystal

Surov and coworker [77] prepared a 1:1 cocrystal of an antiandrogenic drug, bicalutamide (Bic), with salicylamide (2OHBZA) using liquid assisted grinding, slurry sonication as well as solution crystallization. Structural analysis showed that two units of each bicalutamide and salicylamide together form two ring synthon motifs with graph set notation— R 4 4 (12) and R 2 2 (10)—as shown in Figure 9. From the dissolution profile (Figure 10), it was observed that cocrystal Bic with benzamide (BZA) and salicylamide have higher dissolution demonstrating a ‘spring and parachute’ behavior compared to the pure drug bicalutamide.

2.1.4. Caffeine Cocrystal System

Caffeine is one of the most widely used psychoactive drug and central nervous system stimulant used as a model compound to prepare various cocrystals using GRAS coformers as well as APIs. Bucar et al. [78] prepared cocrystals of caffeine with various substituted hydroxybenzoic acid derivatives and studied their synthon preference. Among them, salicylic (2HBA) and gentisic acid (25DHBA) are two coformers of pharmaceutical importance. Cocrystals were discovered using solution mediated phase transformation and showed three different caffine-carboxylic acid heterosynthons (Figure 11 and Figure 12).
Eddleston et al. reported two 1:1 polymorphic cocrystals of caffeine and theophylline in two separate reports [79,80]. The form II caffeine-theophylline cocrystal was prepared using freeze-drying. They proposed that formation of cocrystal proceed via an amorphous phase that is generated as solvent sublimes during the freeze-drying process. Also, they studied the relative stability of the two polymorphic forms and solid–solution phase obtained during freeze-drying and is in the order Form I > Form II > solid-solution, based on theoretical calculation and thermodynamic study. Local arrangement of hydrogen bonded dimers of caffeine and theophylline in the two polymorphic forms are shown in Figure 13.
Nangia et al. reported [81] two multidrug cocrystals of sulfacetamide with caffeine and theophylline along with a few other coformers. Both the cocrystals crystallized in 1:1 ratio having N–H···O and N–H···N synthon between sulfacetamide and coformer (caffeine and theophylline) molecule as shown in Figure 14. Solubility and intrinsic dissolution profile shows that the cocrystals have lower solubility as well as dissolution compared to the pure drug sulfacetamide (Figure 15).
In a very recent report Putra et al. reported [82] a cocrystal of caffeine with epalrestat which is known as an aldose reductase inhibitor for treatment in diabetic neuropathy. The cocrystal showed layered structure that was predicted to be responsible for the higher solubility and dissolution rate of the resultant cocrystal (Figure 16).

2.1.5. Carbamazepine Cocrystal System

Majumder et al. reported [83] a carbamazepine–indomethacin cocrystal using the powder X-ray diffraction (PXRD) structure solution. PXRD data shows a clear formation of cocrystal during ball mill grinding. Crystal structure analysis shows a heterodimer formation between the carboxylic acid group of indomethacin and amide group of carbamazepine molecule as shown in Figure 17.
In another report, Perlovich et al. [84] prepared a 1:1 cocrystal of carbamazepine (CBZ) and 4-aminosalicylic acid (PAS) along with two cocrystal solvates in 2:1:1 ratio of CBZ, PAS and solvent (water and methanol) respectively. From structural analysis it was observed that 1:1 cocrystal of CBZ and PAS contain acid–amide heterosynthon; whereas insertion of solvent molecule results trimeric and tetrameric synthon incorporating solvent molecules water and methanol respectively as shown in Figure 18. From dissolution study, it was observed that cocrystallization of CBZ with PAS decreased the rate of hydrate formation, resulting in a 1.5 times increase in solubility of CBZ.

2.1.6. Dapsone-Drug Cocrystals

Xuefeng Mei and coworker in another report [85], prepared a few drug-drug cocrystals of dapsone. Dapsone is a generic drug used for the treatment of tuberculosis, leprosy, malaria, Kaposi’s sarcoma, dermatoses, and AIDS-related pneumonia. Due to the presence of −NH2 and −SO2− groups, it can form intermolecular hydrogen bond with various coformers. They were able to prepare cocrystals with sulfanilamide, flavone, luteolin, caffeine and 2(3H)-benzothiazolone (Scheme 2). With caffeine, dapsone forms two stoichiometric cocrystals (1:1 and 1:2). Hydrogen bond interaction present in dapsone-caffeine 1:1 cocrystal is shown in Figure 19. Solubility and dissolution rate of cocrystals nicely correlate with coformer solubility order.

2.1.7. Ethenzamide-Gentisic Acid Cocrystals

Aitipamula et al. reported [86] a trimorphic cocrystal of ethenzamide and gentisic acid. Ethanzamide (EA) and gentisic acid (GA) both are nonsteroidal anti-inflammatory drug (NSAID) used mainly in combination with other ingredients for the treatment of mild to moderate pain. During cocrystallization, they obtained three polymorphic forms listed in Table 1.
Looking into the crystal structure, all the three polymorphic forms contain the most stable acid–amide heterosynthon. In the case of form I, the asymmetric unit contains two molecules of each EA and GA. Only one set of the EA and GA sustained by acid–amide heterosynthon (Figure 20). The second symmetry independent GA molecule connects using O–H∙∙∙O hydrogen bond between COOH and 5-OH groups resulting a zipper-like structure.
Form II and form III structure contains one molecule of each EA and GA in asymmetric units. In form II, the acid–amide heterosynthon is further connected via O–H∙∙∙O hydrogen bond of GA molecule resulting a linear tape along the crystallographic b-axis (Figure 21a). In form III, the acid-amide heterosynthon between the EA and GA molecules are further connected via O–H∙∙∙O hydrogen bond forming a tetrameric motif as shown in Figure 21b. Dissolution study shows an enhancement in dissolution rate compared to pure drug EA (Figure 22). Among the three polymorphic forms, metastable polymorphs (form II and form III) have better dissolution compared to form I.

2.1.8. Flufenamic Acid Cocrystals

Trivedi et al. reported [87] two drug-drug cocrystal of flufenamic acid (FFA) with 2-choloro-4-nitrobenzoic acid (CNB) and ethenzamide (EA). Flufenamic acid and EA are non-steroidal anti-inflammatory drugs, whereas CNB is used as a novel potential therapy for immunodeficiency deceases such as an anti-viral and anti-cancer agent. Structural analysis reveals that FFA form acid-acid homosynthon during cocrystallization with CNB, whereas acid-amide heterosynthon was observed in case of FFA–EA cocrystal as shown in Figure 23.
Solubility measurement in 0.1 N HCl (pH = 1) solution shows a five-fold increment in solubility of the cocrystals compared to pure FFA. Also, the synthesized cocrystals were observed to be non-hygroscopic in nature at 75% relative humidity over a period of six months at ambient temperature.
In a different study, Aitipamula and coworker reported [88] a multidrug cocrystal of FFA with theophylline along with two other coformers—namely, 2-pyridone and 4,4′-pyridine. The crystal structure contains theophylline dimer between the saturated N of the imidazole ring and one of the carbonyl group forming N–H···O hydrogen bond. Another O–H···N hydrogen bond involving the carboxylic acid group of FFA and unsaturated N of the imidazole group of theophylline resulting in a four component supramolecular unit as shown in Figure 24a. The resultant cocrystal shows a better solubility and dissolution rate compared to FFA (Figure 24b). The hygroscopic stability was also found to be improved during cocrystallization.
Surov et al. [89] prepared two cocrystals of theophylline with halogen containing APIs named diflunisal and diclofenac having similar hydrogen bond synthon present in FFA-theophylline cocrystal (Figure 25).
The resultant cocrystals showed enhanced apparent solubility; however, intrinsic dissolution rates are comparable to that of the pure drugs.

2.1.9. Furosemide-Caffeine Cocrystals

Goud et al. reported [90] a drug-drug cocrystal of furosemide and caffeine along with few pharmaceutical cocrystals. Furosemide (Lasix) is a loop diuretic drug commonly used for the treatment of hypertension and edema. It is a BCS class IV drug with low solubility (6 mg/L in water) and low permeability (logPow 1.4); whereas caffeine is a central nervous system stimulant and one of the most widely consumed psychoactive drugs. On cocrystallizing furosemide-caffeine 1:1 mixture from a MeOH-MeCN solvent mixture resulted cocrystal of the same in 1:1 stoichiometry. The most acidic COOH group of furosemide forms an O–H···N hydrogen bond with imidazole N of caffeine (Figure 26a). The primary sulfonamide NH donor hydrogen bonds to different caffeine C=O acceptor groups (Figure 26b) and the secondary NH is bonded to one of the sulfonyl O acceptors (Figure 26c).
Solubility analysis showed that furosemide-caffeine cocrystal solubility is six times more compared to pure furosemide. Thus, cocrystals could offer a superior strategy to improve the solubility of BCS class IV drug compared to metastable polymorphs and amorphous dispersions of furosemide.

2.1.10. Gefitinib-Furosemide Salt Hydrate

Thorat et al. reported [91] a salt hydrate of an anticancer drug gefitinib with a loop diuretic drug furosemide. Cocrystallizing stoichiometric amount of gefitinib and furosemide from a 1:1 ethanol-water mixture resulted in gefitinib–furosemide salt hydrate. Carboxylic acid proton of furosemide is transferred to the morpholine ring N of gefitinib molecule. Both APIs interact with each other through strong N–H∙∙∙O hydrogen bond engaging the morpholine group of gefitinib and carboxyl group of furosemide to form API–API dimer (Figure 27). The solubility and dissolution profile shows a solubility order of gefitinib > gefitinib-furosemide salt hydrate > furosemide. Intermediate solubility of molecular salt is due to a stronger hydrogen-bonding synthon, better crystal density, compact crystal packing, and a stable crystal lattice.

2.1.11. Gliclazide-Metformin Salt

Putra et al. reported [92] a multidrug crystal of antidiabetic drug gliclazide (GLI) and metformin (MET). Combining these two non-insulin-dependent diabetes mellitus (NIDDM) drug, they have prepared a salt that is more soluble and have higher dissolution rate compared to GLI; and less hygroscopic compared to MET. The crystal structure contains one molecule of each GLI and MET in the asymmetric unit (Figure 28). A proton from the amide group adjacent to sulfur atom of GLI is transfer to the MET resulting a salt structure. The MET molecules formed a centrosymmetric dimer structure through two N−H···N hydrogen bonds. MET also interacted with three molecules of GLI via relatively strong hydrogen bonds, forming a 1D chain along the a-axis (Figure 29). The hydrophilic MET molecules are sandwiched by hydrophobic GLI molecules, making the salt structure more stable towards hydration.
Dynamic vapor sorption (DVS), solubility, and dissolution rate measurement showed (Figure 30) enhanced physicochemical property of the drug-drug salt compared to the pure components GLI and MET.

2.1.12. Hydrochlorothiazide Cocrystal

Desiraju and coworker reported [93] a few cocrystals of hydrochlorothiazide (HCT) and studied their solubility and membrane diffusivity. During their study, isoniazid an anti-tuberculosis drug was used as a coformer with diuretic drug HCT resulting a cocrystal hydrate in the ratio 1:1:1. Crystal structure contains both sulfonamide dimer as well as amide dimer synthon as shown in Figure 31. In addition the cocrystal structure is stabilized by water with N–H···O and O–H···O hydrogen bond.

2.1.13. Lamivudine-Zidovudine Cocrystal

Desiraju and coworkers reported [94] a drug-drug cocrystal between two anti-HIV drugs, lamivudine and zidovudine. A combination of these two drugs is marketed as a physical mixture with trade name Combivir. On crystallizing the two drug in 1:1 stoichiometry from ethanol resulted a hydrated cocrystal in space group P21 containing a trimeric synthon. The crystal structure contains N–H∙∙∙O and O–H∙∙∙O hydrogen between lamivudine and zidovudine molecule along with a water of crystallization resulting a 1D chain motif as shown in Figure 32. According to their claim, this cocrystal can be a better drug formulation compared to the exiting physical mixture Combivir.

2.1.14. Myricetin–Piracetam Cocrystal

In a recent report by Sowa et. al. [95] prepared a drug–drug cocrystal between myricetin and piracetam using both solvent drop grinding as well as solution crystallization. They found that the piracetam adopt an uncommon conformation not obtained earlier in any of the reported cocrystals. The crystal structure contains N–H···O and O–H···O hydrogen bond between the amide group of piracetam and hydroxyl group of myricetin, resulting in an infinite chain as shown in Figure 33. The resulting cocrystal is stable up to 200 °C.

2.1.15. Norfloxacin–Sulfathiazole Salt Hydrate

In a recent report Desiraju and coworker [96] prepared a new multicomponent solid consisting of antibacterial norfloxacin (NF) and an antimicrobial sulfathiazole (ST) using mechanical grinding. Based on the position of acidic proton of ST the resultant multicomponent solid was confirmed to be a salt hydrate containing a disordered water molecule. From the crystal structure, it was observed that a heterodimer is formed between NF and ST through a proton transfer from the N–H group of ST to the piperazine group of NF. This dimer is further connected via N–H∙∙∙O hydrogen bond between amino group of ST and carboxyl group of NF resulting a sheet-like structure. Successive sheets make channels occupied by disordered water molecules (sustained by O−H···O and C−H···O interactions) along the b-axis (Figure 34). The resultant drug–drug salt hydrate of NF-ST showed enhanced solubility in different pH buffer and improved diffusion rate (Figure 35).
Yeh and Lee in a very recent report [97] prepared two drug-drug cocrystals of ST with theophylline and sulfanilamide using non-conventional method described as intensified crystallization process (Figure 36). The cocrystal ST-sulfanilamide has higher solubility, whereas ST-theophylline has comparable solubility with that of ST crystals.

2.1.16. Oxaprozin-Salbutamol Salt

Aitipamula and coworkers [98] prepared a few multicomponent salts/cocrystals of oxaprozin (OXP) with heteroaromatic compounds—namely piperazine, 4,4′-bipyridine, 1,2-bis(4-pyridyl) ethane along with a drug–drug salt with salbutamol. Oxaprozin (4,5-diphenyl-2-2 oxazole propionic acid) is one of the widely used non-steroidal anti-inflammatory drugs (NSAIDs). OXP has been therapeutically used in inflammatory and painful diseases of rheumatic and nonrheumatic origin, whereas salbutamol (SAL) is an antiasthmatic drug used in the treatment of respiratory disease. In the crystal structure, carboxylic acid proton of OXP is transferred to the secondary amine group of SAL. A four-component supramolecular unit containing a tetrameric synthon of O–H∙∙∙O and N+–H∙∙∙O hydrogen bond results a ladder along b-axis (Figure 37). These ladders are interconnected to each other via C–H∙∙∙O and π∙∙∙π interaction. Surprisingly, this particular salt has a lower solubility and dissolution rate compared to pure SAL and can be a potential formulation for sustained release alternative of pure SAL.

2.1.17. Paracetamol Cocrystal System

Karki et al. [22] carried out mechanical property study of a few cocrystals of acetaminophen—more commonly known as paracetamol (pca)—with various coformers including a few drug molecules namely caffeine, theophylline, theobromine, etc. Cocrystals having layered structures such as paracetamol-theophylline (pca)•(thp) showed better tabletability compared to the existing paracetamol polymorphs (form I and II) (Figure 38). In another report, Lin Lee and Tu Lee [99] prepared caffeine-paracetamol cocrystal by taking p-aminophenol, acetic anhydride, and caffeine on the basis of three component one-pot chemical reaction.

2.1.18. Pyrazinamide-Diflunisal Cocrystal

Eusébio et al. [100] prepared a 1:1 cocrystal of pyrazinamide and diflunisal using neat, liquid-assisted grinding and annealing at 80 °C. Although they were unsuccessful to prepare single crystal, however using spectroscopic and theoretical calculation established the formation of acid-pyridine synthon between diflunisal and pyrazinamide molecule (Figure 39). This particular cocrystal can be used as a combination drug, as pure pyrazinamide has side effects. Moreover, it improves aqueous solubility of diflunisal.

2.1.19. Pyrimethamine-Drug Cocrystals

Jones et al. [101] carried out hydrogen bond propensity calculation (HBPC) in order to predict cocrystal/ salt formation considering pyrimethamine as a model compound. They used various drug and GRAS compounds—namely carbamazepine, theophylline, aspirin, α-keotglutaric acid, saccharin, p-coumaric acid, succinimide, and l-isoleucine. HBPC is useful tool which perform a statistical analysis of the occurrence of hydrogen bonds of the relevant structures present in CSD in order to predict cocrystal/ salt formation. The outcome of their analysis can be summarized in Scheme 3 given below.

2.1.20. Temozolomide Cocrystal System

Nangia et al. reported [102] a few cocrystals of temozolomide, an antitumor prodrug with amide coformers including two pharmaceutically active compounds namely pyrazinamide and caffeine. Hydrogen bond interactions present in the two cocrystal structures are shown in Figure 40.

2.1.21. Theophylline-Acetazolomide Cocrystal

In a very recent report, Kakkar et al. [103] prepared six cocrystal of theophylline and studied their mechanical properties. Among them theophylline-acetazolamide is a drug-drug cocrystal. In the crystal structure, two molecules of each acetazolamide and theophylline form dimers via respective homosynthons. The acetazolamide dimers are further linked with theophylline dimers via N–H∙∙∙O hydrogen bond forming a tape. They are further connected by N–H∙∙∙O and C–H∙∙∙N interactions resulting interlocked 3D network structure (Figure 41a). Due to 3D interlocked structure, theophylline–acetazolamide cocrystal is brittle in nature under mechanical stress (Figure 41b).

2.1.22. Vitamin D2-Vitamin D3 Cocrystal

Xuefeng Mei and coworker reported [104] two cocrystal polymorphs of vitamin D2 (VD2) and D3 (VD3) to improve chemical instability using crystal engineering approach. The two polymorphs, form A (column shaped) and B (plate shaped), were obtained from evaporative crystallization using acetonitrile as solvent. Cocrystal was designed based on hydrogen bond donor–acceptor similarity and to accommodate structural complementarities, selecting α-VD2 and β-VD3 form as starting materials (Figure 42). From structural analysis, it was observed that both the two polymorphs (A and B) have similar hydrogen bond chain structure connecting VD2 and VD3 alternatively in a head-to-head fashion. The only difference between the two structure is overall 3D crystal packing; in form A adjacent polymeric chains are slipped parallel to each other, whereas in form B they are columnar in nature (Figure 43). They have further studied the chemical stability of the two polymorphic cocrystals of VD2 and VD3 stored at 23 °C with an illuminance of 5000 lx and 40 °C/75% RH, and found that form A shows better stability compared to form B and individual compounds (Figure 44).
Zaworotko [105] and Matzger’s [106] group reported a few more multi-drug cocrystal systems of piracetam and lamotrigine using supramolecular heterosynthon. Apart from the examples discussed here, numerous pharmaceutical cocrystals and salts were reported in the literature by some renowned groups—namely Braga [107,108,109,110], Sun [111,112,113,114], Hornedo [9,14,19,34,106,115,116], Myerson [32,33,117], Matzger [106,116,118,119], Bucar [73,78,120,121,122,123,124], and many more [125,126,127,128,129,130,131].

2.2. Drug-Nutraceutical Cocrystals

Nutraceuticals are a recent addition to the coformer family during cocrystal preparation. A nutraceutical can be defined as “a food (or part of a food) that provides medical or health benefits, including the prevention and/or treatment of a disease and possesses a physiological benefit or reduces the risk of chronic disease” [132]. Nutraceutical includes various polyphenols such as phenolic acids, coumarins, stilbenes, flavonoids, various vitamins such as nicotinic acid, retinal, retinoic acid, pyridoxine, l-ascorbic acid, folic acid, menadione, etc. Hydrogen bonding functional group attached to all these molecules facilitates formation of cocrystal with drug molecules resulting pharmaceutical cocrystal. Lawrence and coworkers in a recent review [133] discuss cocrystallization of nutraceuticals as coformers, taking into account physicochemical property modification during drug development. In another review, Caira [134] discusses sulfa drugs as novel cocrystal former during combination therapy, considering their antibacterial and antimicrobial activity.
Schultheiss and coworker mainly focus their research on preparation of various pharmaceutical cocrystal considering nutraceuticals as coformer. In separate reports, they report on cocrystallization of pterostilbene with caffeine and carbamazepine (Figure 45) [135,136]—p-coumaric acid with caffeine, theophylline [137], etc.—along with their solubility enhancement.
Molecular structures of some of the nutraceuticals are shown in Scheme 4.
Nangia et al. also reported [138] few case studies considering nutraceuticals like resorcinol, ferulic acid, vanilic acid, and caffeic acid during cocrystallization with isoniazid. The summary of their result is represented in Figure 46.
Thakuria et al. studied the solubility of various pharmaceutical cocrystals considering nutraceuticals as coformers and correlated coformer solubility with cocrystal solubility of drug molecules viz. ethenzamide [139] and pyrazinamide (Figure 47) [140].
Smith et al. [141] prepare three different cocrystals of quercetin namely quercetin:caffeine (QUECAF), quercetin:caffeine:methanol (QUECAF•MeOH) and quercetin:theobromine:dihydrate (QUETBR•2H2O). Solubility of the QUECAF and QUECAF•MeOH are 14 and 8 times greater whereas QUETBR have slightly enhanced solubility compared to quercetin dihydrate. Dissolution profile of all the cocrystals along with quercetin dihydrate are shown in Figure 48.

3. Future Prospects

From the case studies, it was observed that although multiple reports exist on pharmaceutical cocrystals, drug–drug cocrystals are scarce. The reasons behind the paucity are individual solubility, chemical and physical stability, moisture sensitivity, etc. of APIs. All of these properties can affect the shelf-life of the drugs in multidrug dosage formulation. According to Simon Frantz in his article [142] in Nat. Rev. Drug Discovery, 2006, preparation of multidrug formulation is “a cost centred, not a profit centred” process. Also, they can be used only for a specific set of patients with specific sets of multiple disorders but may not be useful for patients having single ailment; therefore, the importance of drug-drug cocrystal is still debatable. It is needless to mention that preparation of salt formation of a drug containing ionisable sites continues with the objective of improving drug properties rather than cocrystal preparation. However, when routine screening efforts fail to provide desirable crystalline free forms or salts, a possible cocrystal formation between API and a coformer may furnish an appropriate dosage provided a crystalline solid form of that API is a preferred formulation. The outlook is the recent approval of a pharmaceutical cocrystals by the FDA which will continue to gain substantial interest from the pharmaceutical industry as materials for drug development. As a result, R&D departments of many pharmaceutical companies have initiated a fairly large investment in plenty of clinical trials to probe the potential benefits of various coformers/nutraceutical/drug in their cocrystals. Therefore drug-drug and drug-nutraceutical cocrystals and/or salt pre-formulations approach is to identify a suitable dosage formulation to treat targeted disease or multiple disorders simultaneously. It is certain that with more research in this field and reviews [143], they will emerge as a viable alternative to traditional drugs in the near future.

Acknowledgments

Ranjit Thakuria thanks the Department of Science and Technology (DST) for a SERB Young Scientists project (project no. SB/FT/CS-101/2013) for research funding. The Department of Chemistry, Gauhati University is thanked for infrastructure facilities. Bipul Sarma thanks CSIR India ((02(0327)/17/EMR-II) and SERB, India (EMR/2014/000214) for research funding. The Department of Chemical Sciences, Tezpur University is also thanked for infrastructure facilities.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Regulatory Classification of Pharmaceutical Co-Crystals Guidance for Industry. Available online: www.fda.gov/downloads/Drugs/Guidances/UCM516813 (accessed on 2 December 2017).
  2. Reflection paper on the use of cocrystals of active substances in medicinal products. Available online: http://www.ema.europa.eu/docs/en_GB/document_library/Scientific_guideline/2015/07/WC500189927.pdf (accessed on 2 December 2017).
  3. Takagi, T.; Ramachandran, C.; Bermejo, M.; Yamashita, S.; Yu, L.X.; Amidon, G.L. A Provisional Biopharmaceutical Classification of the Top 200 Oral Drug Products in the United States, Great Britain, Spain, and Japan. Mol. Pharm. 2006, 3, 631–643. [Google Scholar] [CrossRef] [PubMed]
  4. Schultheiss, N.; Newman, A. Pharmaceutical Cocrystals and Their Physicochemical Properties. Cryst. Growth Des. 2009, 9, 2950–2967. [Google Scholar] [CrossRef] [PubMed]
  5. Sarma, B.; Saikia, B. Hydrogen bond synthon competition in the stabilization of theophylline cocrystals. CrystEngComm 2014, 16, 4753–4765. [Google Scholar] [CrossRef]
  6. 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] [PubMed]
  7. Saikia, B.; Bora, P.; Khatioda, R.; Sarma, B. Hydrogen Bond Synthons in the Interplay of Solubility and Membrane Permeability/Diffusion in Variable Stoichiometry Drug Cocrystals. Cryst. Growth Des. 2015, 15, 5593–5603. [Google Scholar] [CrossRef]
  8. Saikia, B.; Khatioda, R.; Bora, P.; Sarma, B. Pyridine N-oxides as coformers in the development of drug cocrystals. CrystEngComm 2016, 18, 8454–8464. [Google Scholar] [CrossRef]
  9. Thakuria, R.; Delori, A.; Jones, W.; Lipert, M.P.; Roy, L.; Rodríguez-Hornedo, N. Pharmaceutical cocrystals and poorly soluble drugs. Int. J. Pharm. 2013, 453, 101–125. [Google Scholar] [CrossRef] [PubMed]
  10. Thakuria, R.; Sarma, B.; Nangia, A. 7.03—Hydrogen Bonding in Molecular Crystals A2—Atwood, Jerry L. In Comprehensive Supramolecular Chemistry II; Elsevier: Oxford, UK, 2017; pp. 25–48. [Google Scholar]
  11. Galcera, J.; Molins, E. Effect of the Counterion on the Solubility of Isostructural Pharmaceutical Lamotrigine Salts. Cryst. Growth Des. 2009, 9, 327–334. [Google Scholar] [CrossRef]
  12. Babu, N.J.; Nangia, A. Solubility Advantage of Amorphous Drugs and Pharmaceutical Cocrystals. Cryst. Growth Des. 2011, 11, 2662–2679. [Google Scholar] [CrossRef]
  13. Variankaval, N.; Wenslow, R.; Murry, J.; Hartman, R.; Helmy, R.; Kwong, E.; Clas, S.-D.; Dalton, C.; Santos, I. Preparation and Solid-State Characterization of Nonstoichiometric Cocrystals of a Phosphodiesterase-IV Inhibitor and l-Tartaric Acid. Cryst. Growth Des. 2006, 6, 690–700. [Google Scholar] [CrossRef]
  14. Bethune, S.J.; Huang, N.; Jayasankar, A.; Rodríguez-Hornedo, N. Understanding and Predicting the Effect of Cocrystal Components and pH on Cocrystal Solubility. Cryst. Growth Des. 2009, 9, 3976–3988. [Google Scholar] [CrossRef]
  15. Bhatt, P.M.; Ravindra, N.V.; Banerjee, R.; Desiraju, G.R. Saccharin as a salt former. Enhanced solubilities of saccharinates of active pharmaceutical ingredients. Chem. Commun. 2005, 1073–1075. [Google Scholar] [CrossRef] [PubMed]
  16. Cheney, M.L.; Shan, N.; Healey, E.R.; Hanna, M.; Wojtas, L.; Zaworotko, M.J.; Sava, V.; Song, S.; Sanchez-Ramos, J.R. Effects of Crystal Form on Solubility and Pharmacokinetics: A Crystal Engineering Case Study of Lamotrigine. Cryst. Growth Des. 2010, 10, 394–405. [Google Scholar] [CrossRef]
  17. Aakeröy, C.B.; Forbes, S.; Desper, J. Using Cocrystals to Systematically Modulate Aqueous Solubility and Melting Behavior of an Anticancer Drug. J. Am. Chem. Soc. 2009, 131, 17048–17049. [Google Scholar] [CrossRef] [PubMed]
  18. McNamara, D.P.; Childs, S.L.; Giordano, J.; Iarriccio, A.; Cassidy, J.; Shet, M.S.; Mannion, R.; O’Donnell, E.; Park, A. Use of a Glutaric Acid Cocrystal to Improve Oral Bioavailability of a Low Solubility API. Pharm. Res. 2006, 23, 1888–1897. [Google Scholar] [CrossRef] [PubMed]
  19. Walsh, R.D.B.; Bradner, M.W.; Fleischman, S.; Morales, L.A.; Moulton, B.; Rodriguez-Hornedo, N.; Zaworotko, M.J. Crystal engineering of the composition of pharmaceutical phases. Chem. Commun. 2003, 186–187. [Google Scholar] [CrossRef]
  20. Trask, A.V.; Motherwell, W.D.S.; Jones, W. Pharmaceutical Cocrystallization:  Engineering a Remedy for Caffeine Hydration. Cryst. Growth Des. 2005, 5, 1013–1021. [Google Scholar] [CrossRef]
  21. Trask, A.V.; Motherwell, W.D.S.; Jones, W. Physical stability enhancement of theophylline via cocrystallization. Int. J. Pharm. 2006, 320, 114–123. [Google Scholar] [CrossRef] [PubMed]
  22. Karki, S.; Friščić, T.; Fábián, 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]
  23. Lipinski, C.A. Drug-like properties and the causes of poor solubility and poor permeability. J. Pharmacol. Toxicol. Methods 2000, 44, 235–249. [Google Scholar] [CrossRef]
  24. Thayer, A.M. Finding solutions. Chem. Eng. News 2010, 88, 13–18. [Google Scholar]
  25. Available online: http://apps.who.int/iris/bitstream/10665/43443/1/WHO_TRS_937_eng.pdf (accessed on 2 December 2017).
  26. Hajduk, P.J.; Greer, J. A decade of fragment-based drug design: Strategic advances and lessons learned. Nat. Rev. Drug Discov. 2007, 6, 211–219. [Google Scholar] [CrossRef] [PubMed]
  27. Scott, D.E.; Coyne, A.G.; Hudson, S.A.; Abell, C. Fragment-Based Approaches in Drug Discovery and Chemical Biology. Biochemistry 2012, 51, 4990–5003. [Google Scholar] [CrossRef] [PubMed]
  28. Chellat, M.F.; Raguž, L.; Riedl, R. Targeting Antibiotic Resistance. Angew. Chem. Int. Ed. 2016, 55, 6600–6626. [Google Scholar] [CrossRef] [PubMed]
  29. Sun, W.; Sanderson, P.E.; Zheng, W. Drug combination therapy increases successful drug repositioning. Drug Discov. Today 2016, 21, 1189–1195. [Google Scholar] [CrossRef] [PubMed]
  30. Brittain, H.G. Pharmaceutical cocrystals: The coming wave of new drug substances. J. Pharm. Sci. 2013, 102, 311–317. [Google Scholar] [CrossRef] [PubMed]
  31. Shan, N.; Perry, M.L.; Weyna, D.R.; Zaworotko, M.J. Impact of pharmaceutical cocrystals: The effects on drug pharmacokinetics. Expert Opin. Drug Metab. Toxicol. 2014, 10, 1255–1271. [Google Scholar] [CrossRef] [PubMed]
  32. Chen, J.; Sarma, B.; Evans, J.M.B.; Myerson, A.S. Pharmaceutical Crystallization. Cryst. Growth Des. 2011, 11, 887–895. [Google Scholar] [CrossRef]
  33. Sarma, B.; Chen, J.; Hsi, H.-Y.; Myerson, A.S. Solid forms of pharmaceuticals: Polymorphs, salts and cocrystals. Korean J. Chem. Eng. 2011, 28, 315–322. [Google Scholar] [CrossRef]
  34. Good, D.J.; Rodríguez-Hornedo, N. Solubility Advantage of Pharmaceutical Cocrystals. Cryst. Growth Des. 2009, 9, 2252–2264. [Google Scholar] [CrossRef]
  35. Almarsson, O.; Zaworotko, M.J. Crystal engineering of the composition of pharmaceutical phases. Do pharmaceutical co-crystals represent a new path to improved medicines? Chem. Commun. 2004, 1889–1896. [Google Scholar] [CrossRef] [PubMed]
  36. Generally Regarded as Safe Chemicals by the US-FDA. Available online: https://www.fda.gov/Food/IngredientsPackagingLabeling/FoodAdditivesIngredients/ucm091048.htm (accessed on 2 December 2017).
  37. Lemmerer, A. Covalent assistance to supramolecular synthesis: Modifying the drug functionality of the antituberculosis API isoniazid in situ during co-crystallization with GRAS and API compounds. CrystEngComm 2012, 14, 2465–2478. [Google Scholar] [CrossRef]
  38. Electronic Code of Federal Regulations. Available online: https://www.ecfr.gov/cgi-bin/text-idx?rgn=div5&node=21:3.0.1.1.13 (accessed on 2 December 2017).
  39. Everything Added to Food in the United States (EAFUS). Available online: https://www.accessdata.fda.gov/scripts/fcn/fcnnavigation.cfm?rpt=eafuslisting (accessed on 2 December 2017).
  40. Available online: https://www.pharma.us.novartis.com/sites/www.pharma.us.novartis.com/files/entresto.pdf (accessed on 2 December 2017).
  41. Harrison, W.T.A.; Yathirajan, H.S.; Bindya, S.; Anilkumar, H.G. Escitalopram oxalate: Co-existence of oxalate dianions and oxalic acid molecules in the same crystal. Acta Crystallogr. Sect. C 2007, 63, o129–o131. [Google Scholar] [CrossRef]
  42. Khatioda, R.; Saikia, B.; Das, P.J.; Sarma, B. Solubility and in vitro drug permeation behavior of ethenzamide cocrystals regulated in physiological pH environments. CrystEngComm 2017, 19, 6992–7000. [Google Scholar] [CrossRef]
  43. Thakuria, R.; Cherukuvada, S.; Nangia, A. Crystal Structures of Pyrogallol, Its Hydrate, and Stable Multiple Z′ Cocrystals with N-Heterocycles Containing Metastable Conformers of Pyrogallol. Cryst. Growth Des. 2012, 12, 3944–3953. [Google Scholar] [CrossRef]
  44. Maddileti, D.; Thakuria, R.; Cherukuvada, S.; Nangia, A. Blonanserin HCl salt and its monohydrate. CrystEngComm 2012, 14, 2367–2372. [Google Scholar] [CrossRef]
  45. Kumar, S.S.; Thakuria, R.; Nangia, A. Pharmaceutical cocrystals and a nitrate salt of voriconazole. CrystEngComm 2014, 16, 4722–4731. [Google Scholar] [CrossRef]
  46. Eddleston, M.D.; Thakuria, R.; Aldous, B.J.; Jones, W. An Investigation of the Causes of Cocrystal Dissociation at High Humidity. J. Pharm. Sci. 2014, 103, 2859–2864. [Google Scholar] [CrossRef] [PubMed]
  47. Hoffman, M.; Lindeman, J.A. Chapter 14 Co-Crystals: Commercial Opportunities and Patent Considerations. In Pharmaceutical Salts and Co-Crystals; The Royal Society of Chemistry: London, UK, 2012; pp. 318–329. [Google Scholar]
  48. Buschmann, H.H.D.; Solà, C.L.; Benet, B.J.; Ceron, B.J.C. Co-Crystals of Duloxetine and Co-Crystal Formers for the Treatment of Pain. Patent EP2,123,626, 25 November 2009. [Google Scholar]
  49. Heinrich, B.H.; Farran, J.; Tesson, N. Co-Crystals of Tramadol and Paracetamol. Patent WO2,010,069,561, 24 June 2010. [Google Scholar]
  50. Solá, C.L.; Cerón, B.J.C.; Benet, B.J.; Buschmann, H.H. Co-Crystals of Tramadol and NSAIDs. Patent WO2,010,043,412, 22 April 2010. [Google Scholar]
  51. Plata, S.C.R.; Tesson, N. Co-Crystals of Tramadol and Coxibs. U.S. Patent 8,598,152, 3 December 2013. [Google Scholar]
  52. Plata, S.C.R.; Videla, C.S.; Tesson, N.; Trilla, C.M. Co-Crystals of Venlafaxine and Celecoxib. Patent EP2,515,892, 31 October 2012. [Google Scholar]
  53. Plata, S.C.R.; Tesson, N.; Jiménez, G.C.; Vaiana, L. Crystalline Forms of Sartans Like Telmisartan with Beta Blockers. Patent EP2,649,996, 16 October 2013. [Google Scholar]
  54. Sowa, C.; Gold, R.E.; Chiodo, T.; Vogel, R. Co-Crystals of Cyprodinil and Dithianon. Patent WO2,013,030,777, 7 March 2013. [Google Scholar]
  55. Entresto 97 mg/103 mg Film-Coated Tablets. Available online: http://www.medicines.org.uk/emc/medicine/31244 (accessed on 2 December 2017).
  56. Cafcit. Available online: https://www.drugs.com/pro/cafcit.html (accessed on 3 December 2017).
  57. Smit, J.P.; Hagen, E.J. Polymorphism in Caffeine Citric Acid Cocrystals. J. Chem. Crystallogr. 2015, 45, 128–133. [Google Scholar] [CrossRef]
  58. Karki, S.; Friščić, T.; Jones, W.; Motherwell, W.D.S. Screening for Pharmaceutical Cocrystal Hydrates via Neat and Liquid-Assisted Grinding. Mol. Pharm. 2007, 4, 347–354. [Google Scholar] [CrossRef] [PubMed]
  59. Petruševski, G.; Naumov, P.; Jovanovski, G.; Ng, S.W. Unprecedented sodium–oxygen clusters in the solid-state structure of trisodium hydrogentetravalproate monohydrate: A model for the physiological activity of the anticonvulsant drug Epilim®. Inorg. Chem. Commun. 2008, 11, 81–84. [Google Scholar] [CrossRef]
  60. 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]
  61. Desiraju, G.R. Supramolecular Synthons in Crystal Engineering—A New Organic Synthesis. Angew. Chem. Int. Ed. Engl. 1995, 34, 2311–2327. [Google Scholar] [CrossRef]
  62. Sarma, B.; Bora, P.; Saikia, B. Regulation of π···π Stacking Interactions in Small Molecule Cocrystals and/or Salts for Physiochemical Property Modulation. Cryst. Growth Des. 2018. [Google Scholar] [CrossRef]
  63. Bhogala, B.R.; Basavoju, S.; Nangia, A. Three-Component Carboxylic Acid−Bipyridine Lattice Inclusion Host. Supramolecular Synthesis of Ternary Cocrystals. Cryst. Growth Des. 2005, 5, 1683–1686. [Google Scholar] [CrossRef]
  64. Sarma, B.; Nath, N.K.; Bhogala, B.R.; Nangia, A. Synthon Competition and Cooperation in Molecular Salts of Hydroxybenzoic Acids and Aminopyridines. Cryst. Growth Des. 2009, 9, 1546–1557. [Google Scholar] [CrossRef]
  65. Bhogala, B.R.; Nangia, A. Cocrystals of 1,3,5-Cyclohexanetricarboxylic Acid with 4,4′-Bipyridine Homologues:  Acid···Pyridine Hydrogen Bonding in Neutral and Ionic Complexes. Cryst. Growth Des. 2003, 3, 547–554. [Google Scholar] [CrossRef]
  66. 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]
  67. Cruz-Cabeza, A.J. Acid-base crystalline complexes and the pKa rule. CrystEngComm 2012, 14, 6362–6365. [Google Scholar] [CrossRef]
  68. Sarmah, K.K.; Sarma, A.; Roy, K.; Rao, D.R.; Thakuria, R. Olanzapine Salts and Diversity in Molecular Packing. Cryst. Growth Des. 2016, 16, 1047–1055. [Google Scholar] [CrossRef]
  69. Thakuria, R.; Nangia, A. Olanzapinium Salts, Isostructural Solvates, and Their Physicochemical Properties. Cryst. Growth Des. 2013, 13, 3672–3680. [Google Scholar] [CrossRef]
  70. Thakuria, R.; Nangia, A. Highly soluble olanzapinium maleate crystalline salts. CrystEngComm 2011, 13, 1759–1764. [Google Scholar] [CrossRef]
  71. Sarma, B.; Thakuria, R.; Nath, N.K.; Nangia, A. Crystal structures of mirtazapine molecular salts. CrystEngComm 2011, 13, 3232–3240. [Google Scholar] [CrossRef]
  72. Laurence, C.; Berthelot, M. Observations on the strength of hydrogen bonding. Perspect. Drug Discov. Des. 2000, 18, 39–60. [Google Scholar] [CrossRef]
  73. Zegarac, M.; Leksic, E.; Sket, P.; Plavec, J.; Devcic Bogdanovic, M.; Bucar, D.-K.; Dumic, M.; Mestrovic, E. A sildenafil cocrystal based on acetylsalicylic acid exhibits an enhanced intrinsic dissolution rate. CrystEngComm 2014, 16, 32–35. [Google Scholar] [CrossRef]
  74. 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]
  75. Grobelny, P.; Mukherjee, A.; Desiraju, G.R. Drug-drug co-crystals: Temperature-dependent proton mobility in the molecular complex of isoniazid with 4-aminosalicylic acid. CrystEngComm 2011, 13, 4358–4364. [Google Scholar] [CrossRef]
  76. Liu, F.; Song, Y.; Liu, Y.-N.; Li, Y.-T.; Wu, Z.-Y.; Yan, C.-W. Drug-bridge-drug ternary cocrystallization strategy for anti-tuberculosis drugs combination. Cryst. Growth Des. 2018. [Google Scholar] [CrossRef]
  77. Surov, A.O.; Solanko, K.A.; Bond, A.D.; Bauer-Brandl, A.; Perlovich, G.L. Cocrystals of the antiandrogenic drug bicalutamide: Screening, crystal structures, formation thermodynamics and lattice energies. CrystEngComm 2016, 18, 4818–4829. [Google Scholar] [CrossRef]
  78. Bučar, D.-K.; Henry, R.F.; Lou, X.; Duerst, R.W.; MacGillivray, L.R.; Zhang, G.G.Z. Cocrystals of Caffeine and Hydroxybenzoic Acids Composed of Multiple Supramolecular Heterosynthons: Screening via Solution-Mediated Phase Transformation and Structural Characterization. Cryst. Growth Des. 2009, 9, 1932–1943. [Google Scholar] [CrossRef]
  79. Eddleston, M.D.; Lloyd, G.O.; Jones, W. Cocrystal dissociation and molecular demixing in the solid state. Chem. Commun. 2012, 48, 8075–8077. [Google Scholar] [CrossRef] [PubMed]
  80. Eddleston, M.D.; Patel, B.; Day, G.M.; Jones, W. Cocrystallization by Freeze-Drying: Preparation of Novel Multicomponent Crystal Forms. Cryst. Growth Des. 2013, 13, 4599–4606. [Google Scholar] [CrossRef]
  81. Rajesh Goud, N.; Khan, R.A.; Nangia, A. Modulating the solubility of sulfacetamide by means of cocrystals. CrystEngComm 2014, 16, 5859–5869. [Google Scholar] [CrossRef]
  82. Putra, O.D.; Umeda, D.; Nugraha, Y.P.; Furuishi, T.; Nagase, H.; Fukuzawa, K.; Uekusa, H.; Yonemochi, E. Solubility improvement of epalrestat by layered structure formation via cocrystallization. CrystEngComm 2017, 19, 2614–2622. [Google Scholar] [CrossRef]
  83. Majumder, M.; Buckton, G.; Rawlinson-Malone, C.; Williams, A.C.; Spillman, M.J.; Shankland, N.; Shankland, K. A carbamazepine-indomethacin (1:1) cocrystal produced by milling. CrystEngComm 2011, 13, 6327–6328. [Google Scholar] [CrossRef]
  84. Drozd, K.V.; Manin, A.N.; Churakov, A.V.; Perlovich, G.L. Novel drug-drug cocrystals of carbamazepine with para-aminosalicylic acid: Screening, crystal structures and comparative study of carbamazepine cocrystal formation thermodynamics. CrystEngComm 2017, 19, 4273–4286. [Google Scholar] [CrossRef]
  85. Jiang, L.; Huang, Y.; Zhang, Q.; He, H.; Xu, Y.; Mei, X. Preparation and Solid-State Characterization of Dapsone Drug–Drug Co-Crystals. Cryst. Growth Des. 2014, 14, 4562–4573. [Google Scholar] [CrossRef]
  86. Aitipamula, S.; Chow, P.S.; Tan, R.B.H. Trimorphs of a pharmaceutical cocrystal involving two active pharmaceutical ingredients: Potential relevance to combination drugs. CrystEngComm 2009, 11, 1823–1827. [Google Scholar] [CrossRef]
  87. Nechipadappu, S.K.; Tekuri, V.; Trivedi, D.R. Pharmaceutical Co-Crystal of Flufenamic Acid: Synthesis and Characterization of Two Novel Drug-Drug Co-Crystal. J. Pharm. Sci. 2017, 106, 1384–1390. [Google Scholar] [CrossRef] [PubMed]
  88. 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]
  89. 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] [PubMed]
  90. Goud, N.R.; Gangavaram, S.; Suresh, K.; Pal, S.; Manjunatha, S.G.; Nambiar, S.; Nangia, A. Novel furosemide cocrystals and selection of high solubility drug forms. J. Pharm. Sci. 2012, 101, 664–680. [Google Scholar] [CrossRef] [PubMed]
  91. Thorat, S.H.; Sahu, S.K.; Patwadkar, M.V.; Badiger, M.V.; Gonnade, R.G. Drug–Drug Molecular Salt Hydrate of an Anticancer Drug Gefitinib and a Loop Diuretic Drug Furosemide: An Alternative for Multidrug Treatment. J. Pharm. Sci. 2015, 104, 4207–4216. [Google Scholar] [CrossRef] [PubMed]
  92. Putra, O.D.; Furuishi, T.; Yonemochi, E.; Terada, K.; Uekusa, H. Drug–Drug Multicomponent Crystals as an Effective Technique to Overcome Weaknesses in Parent Drugs. Cryst. Growth Des. 2016, 16, 3577–3581. [Google Scholar] [CrossRef]
  93. Gopi, S.P.; Banik, M.; Desiraju, G.R. New Cocrystals of Hydrochlorothiazide: Optimizing Solubility and Membrane Diffusivity. Cryst. Growth Des. 2017, 17, 308–316. [Google Scholar] [CrossRef]
  94. 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]
  95. Sowa, M.; Ślepokura, K.; Matczak-Jon, E. A 1:1 pharmaceutical cocrystal of myricetin in combination with uncommon piracetam conformer: X-ray single crystal analysis and mechanochemical synthesis. J. Mol. Struct. 2014, 1058, 114–121. [Google Scholar] [CrossRef]
  96. Gopi, S.P.; Ganguly, S.; Desiraju, G.R. A Drug–Drug Salt Hydrate of Norfloxacin and Sulfathiazole: Enhancement of in Vitro Biological Properties via Improved Physicochemical Properties. Mol. Pharm. 2016, 13, 3590–3594. [Google Scholar] [CrossRef] [PubMed]
  97. Yeh, K.L.; Lee, T. Intensified Crystallization Processes for 1:1 Drug-Drug Co-crystals of Sulfathiazole-Theophylline, and Sulfathiazole-Sulfanilamide. Cryst. Growth Des. 2018. [Google Scholar] [CrossRef]
  98. Aitipamula, S.; Wong, A.B.H.; Chow, P.S.; Tan, R.B.H. Novel solid forms of oxaprozin: Cocrystals and an extended release drug-drug salt of salbutamol. RSC Adv. 2016, 6, 34110–34119. [Google Scholar] [CrossRef]
  99. Lee, H.L.; Lee, T. Direct co-crystal assembly from synthesis to co-crystallization. CrystEngComm 2015, 17, 9002–9006. [Google Scholar] [CrossRef]
  100. Évora, A.O.L.; Castro, R.A.E.; Maria, T.M.R.; Rosado, M.T.S.; Ramos Silva, M.; Matos Beja, A.; Canotilho, J.; Eusébio, M.E.S. Pyrazinamide-Diflunisal: A New Dual-Drug Co-Crystal. Cryst. Growth Des. 2011, 11, 4780–4788. [Google Scholar] [CrossRef]
  101. Delori, A.; Galek, P.T.A.; Pidcock, E.; Patni, M.; Jones, W. Knowledge-based hydrogen bond prediction and the synthesis of salts and cocrystals of the anti-malarial drug pyrimethamine with various drug and GRAS molecules. CrystEngComm 2013, 15, 2916–2928. [Google Scholar] [CrossRef]
  102. Sanphui, P.; Babu, N.J.; Nangia, A. Temozolomide Cocrystals with Carboxamide Coformers. Cryst. Growth Des. 2013, 13, 2208–2219. [Google Scholar] [CrossRef]
  103. Kakkar, S.; Bhattacharya, B.; Reddy, C.M.; Ghosh, S. Tuning mechanical behaviour by controlling the structure of a series of theophylline co-crystals. CrystEngComm 2018. [Google Scholar] [CrossRef]
  104. Wang, J.-R.; Yu, Q.; Dai, W.; Mei, X. Drug-drug co-crystallization presents a new opportunity for the development of stable vitamins. Chem. Commun. 2016, 52, 3572–3575. [Google Scholar] [CrossRef] [PubMed]
  105. Vishweshwar, P.; McMahon, J.A.; Peterson, M.L.; Hickey, M.B.; Shattock, T.R.; Zaworotko, M.J. Crystal engineering of pharmaceutical co-crystals from polymorphic active pharmaceutical ingredients. Chem. Commun. 2005, 4601–4603. [Google Scholar] [CrossRef] [PubMed]
  106. Kaur, R.; Cavanagh, K.L.; Rodríguez-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]
  107. Braga, D.; Grepioni, F.; Chelazzi, L.; Campana, M.; Confortini, D.; Viscomi, G.C. The structure-property relationship of four crystal forms of rifaximin. CrystEngComm 2012, 14, 6404–6411. [Google Scholar] [CrossRef]
  108. Braga, D.; Grepioni, F.; Maini, L.; Capucci, D.; Nanna, S.; Wouters, J.; Aerts, L.; Quere, L. Combining piracetam and lithium salts: Ionic co-crystals and co-drugs? Chem. Commun. 2012, 48, 8219–8221. [Google Scholar] [CrossRef] [PubMed]
  109. Grifasi, F.; Chierotti, M.R.; Gaglioti, K.; Gobetto, R.; Maini, L.; Braga, D.; Dichiarante, E.; Curzi, M. Using Salt Cocrystals to Improve the Solubility of Niclosamide. Cryst. Growth Des. 2015, 15, 1939–1948. [Google Scholar] [CrossRef]
  110. Braga, D.; Chelazzi, L.; Grepioni, F.; Dichiarante, E.; Chierotti, M.R.; Gobetto, R. Molecular Salts of Anesthetic Lidocaine with Dicarboxylic Acids: Solid-State Properties and a Combined Structural and Spectroscopic Study. Cryst. Growth Des. 2013, 13, 2564–2572. [Google Scholar] [CrossRef]
  111. 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]
  112. Chow, S.F.; Chen, M.; Shi, L.; Chow, A.H.L.; 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] [PubMed]
  113. Sun, C.C. Cocrystallization for successful drug delivery. Expert Opin. Drug Deliv. 2013, 10, 201–213. [Google Scholar] [CrossRef] [PubMed]
  114. 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]
  115. Jayasankar, A.; Reddy, L.S.; Bethune, S.J.; Rodríguez-Hornedo, N. Role of Cocrystal and Solution Chemistry on the Formation and Stability of Cocrystals with Different Stoichiometry. Cryst. Growth Des. 2009, 9, 889–897. [Google Scholar] [CrossRef]
  116. Rodríguez-Spong, B.; Price, C.P.; Jayasankar, A.; Matzger, A.J.; Rodrı́guez-Hornedo, N.R. General principles of pharmaceutical solid polymorphism: A supramolecular perspective. Adv. Drug Deliv. Rev. 2004, 56, 241–274. [Google Scholar] [CrossRef] [PubMed]
  117. Wang, M.; Rutledge, G.C.; Myerson, A.S.; Trout, B.L. Production and Characterization of Carbamazepine Nanocrystals by Electrospraying for Continuous Pharmaceutical Manufacturing. J. Pharm. Sci. 2012, 101, 1178–1188. [Google Scholar] [CrossRef] [PubMed]
  118. Li, Z.; Matzger, A.J. Influence of Coformer Stoichiometric Ratio on Pharmaceutical Cocrystal Dissolution: Three Cocrystals of Carbamazepine/4-Aminobenzoic Acid. Mol. Pharm. 2016, 13, 990–995. [Google Scholar] [CrossRef] [PubMed]
  119. Porter Iii, W.W.; Elie, S.C.; Matzger, A.J. Polymorphism in Carbamazepine Cocrystals. Cryst. Growth Des. 2008, 8, 14–16. [Google Scholar] [CrossRef] [PubMed]
  120. Bučar, D.-K.; Elliott, J.A.; Eddleston, M.D.; Cockcroft, J.K.; Jones, W. Sonocrystallization Yields Monoclinic Paracetamol with Significantly Improved Compaction Behavior. Angew. Chem. 2015, 127, 251–255. [Google Scholar] [CrossRef]
  121. Bucar, D.-K.; Day, G.M.; Halasz, I.; Zhang, G.G.Z.; Sander, J.R.G.; Reid, D.G.; MacGillivray, L.R.; Duer, M.J.; Jones, W. The curious case of (caffeine)[middle dot](benzoic acid): How heteronuclear seeding allowed the formation of an elusive cocrystal. Chem. Sci. 2013, 4, 4417–4425. [Google Scholar] [CrossRef]
  122. Bucar, D.-K.; Henry, R.F.; Lou, X.; Borchardt, T.B.; Zhang, G.G.Z. A “hidden” co-crystal of caffeine and adipic acid. Chem. Commun. 2007, 525–527. [Google Scholar] [CrossRef] [PubMed]
  123. Sander, J.R.G.; Bučar, D.-K.; Henry, R.F.; Baltrusaitis, J.; Zhang, G.G.Z.; Macgillivray, L.R. A Red Zwitterionic Co-Crystal of Acetaminophen and 2,4-Pyridinedicarboxylic Acid. J. Pharm. Sci. 2010, 99, 3676–3683. [Google Scholar] [CrossRef] [PubMed]
  124. Bučar, D.-K.; Henry, R.F.; Duerst, R.W.; Lou, X.; MacGillivray, L.R.; Zhang, G.G.Z. A 1:1 Cocrystal of Caffeine and 2-Hydroxy-1-Naphthoic Acid Obtained via a Slurry Screening Method. J. Chem. Crystallogr. 2010, 40, 933–939. [Google Scholar] [CrossRef]
  125. Berry, D.J.; Seaton, C.C.; Clegg, W.; Harrington, R.W.; Coles, S.J.; Horton, P.N.; Hursthouse, M.B.; Storey, R.; Jones, W.; Friščić, T.; et al. Applying Hot-Stage Microscopy to Co-Crystal Screening: A Study of Nicotinamide with Seven Active Pharmaceutical Ingredients. Cryst. Growth Des. 2008, 8, 1697–1712. [Google Scholar] [CrossRef]
  126. Velaga, S.P.; Basavoju, S.; Boström, D. Norfloxacin saccharinate–saccharin dihydrate cocrystal—A new pharmaceutical cocrystal with an organic counter ion. J. Mol. Struct. 2008, 889, 150–153. [Google Scholar] [CrossRef]
  127. Hong, C.; Xie, Y.; Yao, Y.; Li, G.; Yuan, X.; Shen, H. A Novel Strategy for Pharmaceutical Cocrystal Generation without Knowledge of Stoichiometric Ratio: Myricetin Cocrystals and a Ternary Phase Diagram. Pharm. Res. 2015, 32, 47–60. [Google Scholar] [CrossRef] [PubMed]
  128. Basavoju, S.; Boström, D.; Velaga, S.P. Indomethacin–Saccharin Cocrystal: Design, Synthesis and Preliminary Pharmaceutical Characterization. Pharm. Res. 2008, 25, 530–541. [Google Scholar] [CrossRef] [PubMed]
  129. Weyna, D.R.; Shattock, T.; Vishweshwar, P.; Zaworotko, M.J. Synthesis and Structural Characterization of Cocrystals and Pharmaceutical Cocrystals: Mechanochemistry vs Slow Evaporation from Solution. Cryst. Growth Des. 2009, 9, 1106–1123. [Google Scholar] [CrossRef]
  130. Basavoju, S.; Boström, D.; Velaga, S.P. Pharmaceutical Cocrystal and Salts of Norfloxacin. Cryst. Growth Des. 2006, 6, 2699–2708. [Google Scholar] [CrossRef]
  131. Friščić, T.; Jones, W. Benefits of cocrystallisation in pharmaceutical materials science: An update. J. Pharm. Pharmacol. 2010, 62, 1547–1559. [Google Scholar] [CrossRef] [PubMed]
  132. Kalra, E.K. Nutraceutical-definition and introduction. AAPS PharmSci 2003, 5, 27–28. [Google Scholar] [CrossRef] [PubMed]
  133. Sinha, A.S.; Maguire, A.R.; Lawrence, S.E. Cocrystallization of Nutraceuticals. Cryst. Growth Des. 2015, 15, 984–1009. [Google Scholar] [CrossRef]
  134. Caira, M.R. Sulfa Drugs as Model Cocrystal Formers. Mol. Pharm. 2007, 4, 310–316. [Google Scholar] [CrossRef] [PubMed]
  135. Bethune, S.J.; Schultheiss, N.; Henck, J.-O. Improving the Poor Aqueous Solubility of Nutraceutical Compound Pterostilbene through Cocrystal Formation. Cryst. Growth Des. 2011, 11, 2817–2823. [Google Scholar] [CrossRef]
  136. Schultheiss, N.; Bethune, S.; Henck, J.-O. Nutraceutical cocrystals: Utilizing pterostilbene as a cocrystal former. CrystEngComm 2010, 12, 2436–2442. [Google Scholar] [CrossRef]
  137. Schultheiss, N.; Roe, M.; Boerrigter, S.X.M. Cocrystals of nutraceutical p-coumaric acid with caffeine and theophylline: Polymorphism and solid-state stability explored in detail using their crystal graphs. CrystEngComm 2011, 13, 611–619. [Google Scholar] [CrossRef]
  138. Swapna, B.; Maddileti, D.; Nangia, A. Cocrystals of the Tuberculosis Drug Isoniazid: Polymorphism, Isostructurality, and Stability. Cryst. Growth Des. 2014, 14, 5991–6005. [Google Scholar] [CrossRef]
  139. Sarmah, K.K.; Boro, K.; Arhangelskis, M.; Thakuria, R. Crystal structure landscape of ethenzamide: A physicochemical property study. CrystEngComm 2017, 19, 826–833. [Google Scholar] [CrossRef]
  140. Sarmah, K.K.; Rajbongshi, T.; Bhowmick, S.; Thakuria, R. First-line antituberculosis drug, pyrazinamide, its pharmaceutically relevant cocrystals and a salt. Acta Crystallogr. Sect. B 2017, 73, 1007–1016. [Google Scholar] [CrossRef] [PubMed]
  141. Smith, A.J.; Kavuru, P.; Wojtas, L.; Zaworotko, M.J.; Shytle, R.D. Cocrystals of Quercetin with Improved Solubility and Oral Bioavailability. Mol. Pharm. 2011, 8, 1867–1876. [Google Scholar] [CrossRef] [PubMed]
  142. Simon, F. The trouble with making combination drugs. Nat. Rev. Drug Discov. 2006, 5, 881–882. [Google Scholar] [CrossRef] [PubMed]
  143. Thipparaboina, R.; Kumar, D.; Chavan, R.B.; Shastri, N.R. Multidrug co-crystals: Towards the development of effective therapeutic hybrids. Drug Discov. Today 2016, 21, 481–490. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Typical representation of preparation of cocrystal.
Figure 1. Typical representation of preparation of cocrystal.
Crystals 08 00101 g001
Figure 2. The ‘spring and parachute’ approach to promoting and maintaining supersaturation of poorly soluble drugs. Adapted with permission from Cryst. Growth Des., 2011, 11 (7), 2662–2679. Copyright 2011 American Chemical Society.
Figure 2. The ‘spring and parachute’ approach to promoting and maintaining supersaturation of poorly soluble drugs. Adapted with permission from Cryst. Growth Des., 2011, 11 (7), 2662–2679. Copyright 2011 American Chemical Society.
Crystals 08 00101 g002
Scheme 1. (a) The pKa rules of thumb to predict the H-bonding motifs in multi-component crystals. This Scheme is reproduced with a slender modification from Ph.D. thesis entitled Structural and Thermal Analysis of Organic Solids by B. Sarma, 2009, University of Hyderabad, India. (b) Calculated occurrence of AB (grey) and AB+ (orange) as a function of the calculated ∆pKa Inculcate from a CSD study of a set of 6465 acid-base complexes carried out by Aurora; Adapted with permission from CrystEngComm, 2012, 14, 6362–6365. Copyright 2012 the Royal Society of Chemistry.
Scheme 1. (a) The pKa rules of thumb to predict the H-bonding motifs in multi-component crystals. This Scheme is reproduced with a slender modification from Ph.D. thesis entitled Structural and Thermal Analysis of Organic Solids by B. Sarma, 2009, University of Hyderabad, India. (b) Calculated occurrence of AB (grey) and AB+ (orange) as a function of the calculated ∆pKa Inculcate from a CSD study of a set of 6465 acid-base complexes carried out by Aurora; Adapted with permission from CrystEngComm, 2012, 14, 6362–6365. Copyright 2012 the Royal Society of Chemistry.
Crystals 08 00101 sch001
Figure 3. Dissolution rates of 1:1 sildenafil–acetylsalicylic acid cocrystal and sildenafil citrate in (a) an aqueous solution at a pH of 1.2, (b) degassed water, and (c) an aqueous solution containing 1.2% NaCl. Colour scheme: 1:1 cocrystal—blue, sildenafil citrate—red. Adapted with permission from CrystEngComm, 2014, 16, 32-35. Copyright 2014 The Royal Society of Chemistry.
Figure 3. Dissolution rates of 1:1 sildenafil–acetylsalicylic acid cocrystal and sildenafil citrate in (a) an aqueous solution at a pH of 1.2, (b) degassed water, and (c) an aqueous solution containing 1.2% NaCl. Colour scheme: 1:1 cocrystal—blue, sildenafil citrate—red. Adapted with permission from CrystEngComm, 2014, 16, 32-35. Copyright 2014 The Royal Society of Chemistry.
Crystals 08 00101 g003
Figure 4. Supramolecular synthon present in 1:1 cocrystal of meloxicam and aspirin.
Figure 4. Supramolecular synthon present in 1:1 cocrystal of meloxicam and aspirin.
Crystals 08 00101 g004
Figure 5. Hydrogen bond interaction present in INH-PAS cocrystal showing finite tetramer motif highlighted in green. Adapted with permission from CrystEngComm, 2011, 13, 4358–4364. Copyright 2011, The Royal Society of Chemistry.
Figure 5. Hydrogen bond interaction present in INH-PAS cocrystal showing finite tetramer motif highlighted in green. Adapted with permission from CrystEngComm, 2011, 13, 4358–4364. Copyright 2011, The Royal Society of Chemistry.
Crystals 08 00101 g005
Figure 6. COOH···Npyridine and amide-amide dimer synthon between the symmetry independent molecules of INH-PAS cocrystal. Adapted with permission from CrystEngComm, 2011, 13, 4358–4364. Copyright 2011, The Royal Society of Chemistry.
Figure 6. COOH···Npyridine and amide-amide dimer synthon between the symmetry independent molecules of INH-PAS cocrystal. Adapted with permission from CrystEngComm, 2011, 13, 4358–4364. Copyright 2011, The Royal Society of Chemistry.
Crystals 08 00101 g006
Figure 7. Hydrogen bond interaction present in Pyz-PAS-H2O cocrystal. Adapted with permission from CrystEngComm, 2011, 13, 4358–4364. Copyright 2011, The Royal Society of Chemistry.
Figure 7. Hydrogen bond interaction present in Pyz-PAS-H2O cocrystal. Adapted with permission from CrystEngComm, 2011, 13, 4358–4364. Copyright 2011, The Royal Society of Chemistry.
Crystals 08 00101 g007
Figure 8. Molecular structures of INH, Pyz, and FA; crystal structure of INH-FA-Pyz ternary cocrystal. Adapted with permission from Cryst. Growth Des., DOI: 10.1021/acs.cgd.7b01738. Copyright 2018, American Chemical Society.
Figure 8. Molecular structures of INH, Pyz, and FA; crystal structure of INH-FA-Pyz ternary cocrystal. Adapted with permission from Cryst. Growth Des., DOI: 10.1021/acs.cgd.7b01738. Copyright 2018, American Chemical Society.
Crystals 08 00101 g008
Figure 9. Two ring synthon motifs with graph set notation R 4 4 (12) and R 2 2 (10) between the tetrameric unit of a bicalutamide–salicylamide cocrystal. Adapted with permission from CrystEngComm, 2016, 18, 4818–4829. Copyright 2016, The Royal Society of Chemistry.
Figure 9. Two ring synthon motifs with graph set notation R 4 4 (12) and R 2 2 (10) between the tetrameric unit of a bicalutamide–salicylamide cocrystal. Adapted with permission from CrystEngComm, 2016, 18, 4818–4829. Copyright 2016, The Royal Society of Chemistry.
Crystals 08 00101 g009
Figure 10. Dissolution profiles for the cocrystals (Bic + BZA and Bic + 2OHBZA) and pure bicalutamide in the pH 7.4 phosphate buffer at 25 °C. Adapted with permission from CrystEngComm, 2016, 18, 4818–4829. Copyright 2016, The Royal Society of Chemistry.
Figure 10. Dissolution profiles for the cocrystals (Bic + BZA and Bic + 2OHBZA) and pure bicalutamide in the pH 7.4 phosphate buffer at 25 °C. Adapted with permission from CrystEngComm, 2016, 18, 4818–4829. Copyright 2016, The Royal Society of Chemistry.
Crystals 08 00101 g010
Figure 11. (a) Acid-imidazole synthons between 2HBA and caffeine, respectively. (bd) Crystal packing of the resultant 1:1 cocrystal. Adapted with permission from Cryst. Growth Des., 2009, 9 (4), 1932–1943. Copyright 2009, American Chemical Society.
Figure 11. (a) Acid-imidazole synthons between 2HBA and caffeine, respectively. (bd) Crystal packing of the resultant 1:1 cocrystal. Adapted with permission from Cryst. Growth Des., 2009, 9 (4), 1932–1943. Copyright 2009, American Chemical Society.
Crystals 08 00101 g011
Figure 12. (a) Acid-imidazole dimer and hydroxyl–amide dimer between 25DHBA and caffeine resulting a 1D polymeric chain. (b,c) Stacked packing of the 1D polymer sustained by weak C–H···O hydrogen bond. Adapted with permission from Cryst. Growth Des., 2009, 9 (4), 1932–1943. Copyright 2009, American Chemical Society.
Figure 12. (a) Acid-imidazole dimer and hydroxyl–amide dimer between 25DHBA and caffeine resulting a 1D polymeric chain. (b,c) Stacked packing of the 1D polymer sustained by weak C–H···O hydrogen bond. Adapted with permission from Cryst. Growth Des., 2009, 9 (4), 1932–1943. Copyright 2009, American Chemical Society.
Crystals 08 00101 g012
Figure 13. Hydrogen bonded dimer of caffeine and theophylline in two respective caffine-theophylline cocrystal polymorphs. Adapted with permission from Cryst. Growth Des., 2013, 13 (10), 4599–4606. Copyright 2013, American Chemical Society.
Figure 13. Hydrogen bonded dimer of caffeine and theophylline in two respective caffine-theophylline cocrystal polymorphs. Adapted with permission from Cryst. Growth Des., 2013, 13 (10), 4599–4606. Copyright 2013, American Chemical Society.
Crystals 08 00101 g013
Figure 14. N–H···O and N–H···N hydrogen bond connecting sulfacetamide and coformer (a) caffeine and (b) theophylline in the respective cocrystals. Adapted with permission from CrystEngComm, 2014, 16, 5859–5869. Copyright 2014, The Royal Society of Chemistry.
Figure 14. N–H···O and N–H···N hydrogen bond connecting sulfacetamide and coformer (a) caffeine and (b) theophylline in the respective cocrystals. Adapted with permission from CrystEngComm, 2014, 16, 5859–5869. Copyright 2014, The Royal Society of Chemistry.
Crystals 08 00101 g014
Figure 15. Intrinsic dissolution of sulfacetamide cocrystals in pH = 7 phosphate buffer. Adapted with permission from CrystEngComm, 2014, 16, 5859–5869. Copyright 2014, The Royal Society of Chemistry.
Figure 15. Intrinsic dissolution of sulfacetamide cocrystals in pH = 7 phosphate buffer. Adapted with permission from CrystEngComm, 2014, 16, 5859–5869. Copyright 2014, The Royal Society of Chemistry.
Crystals 08 00101 g015
Figure 16. (a) Layered structure of 1:1 caffeine-epalrestat cocrystal connected through O–H···O hydrogen bond. Hydrogen atoms were removed for clarity. (b) Intrinsic dissolution rate of caffeine-epalrestat cocrystal (red) compared with pure epalrestat drug (blue). Adapted with permission from CrystEngComm, 2017, 19, 2614–2622. Copyright 2017, The Royal Society of Chemistry.
Figure 16. (a) Layered structure of 1:1 caffeine-epalrestat cocrystal connected through O–H···O hydrogen bond. Hydrogen atoms were removed for clarity. (b) Intrinsic dissolution rate of caffeine-epalrestat cocrystal (red) compared with pure epalrestat drug (blue). Adapted with permission from CrystEngComm, 2017, 19, 2614–2622. Copyright 2017, The Royal Society of Chemistry.
Crystals 08 00101 g016
Figure 17. Acid-amide dimer synthon present in carbamazepine-indomethacin cocrystal obtained using PXRD structure solution. Adapted with permission from CrystEngComm, 2011, 13, 6327–6328. Copyright 2011, The Royal Society of Chemistry.
Figure 17. Acid-amide dimer synthon present in carbamazepine-indomethacin cocrystal obtained using PXRD structure solution. Adapted with permission from CrystEngComm, 2011, 13, 6327–6328. Copyright 2011, The Royal Society of Chemistry.
Crystals 08 00101 g017
Figure 18. Hydrogen bond interaction present in (a) 1:1 CBZ-PAS cocrystal; (b) 2:1:1 CBZ-PAS-water and (c) 2:1:1 CBZ-PAS-methanol structure. Adapted with permission from CrystEngComm, 2017, 19, 4273–4286. Copyright 2017, The Royal Society of Chemistry.
Figure 18. Hydrogen bond interaction present in (a) 1:1 CBZ-PAS cocrystal; (b) 2:1:1 CBZ-PAS-water and (c) 2:1:1 CBZ-PAS-methanol structure. Adapted with permission from CrystEngComm, 2017, 19, 4273–4286. Copyright 2017, The Royal Society of Chemistry.
Crystals 08 00101 g018
Scheme 2. Molecular structure of Dapsone and coformers studied.
Scheme 2. Molecular structure of Dapsone and coformers studied.
Crystals 08 00101 sch002
Figure 19. Hydrogen bond interaction present in dapsone-caffeine 1:1 cocrystal. Adapted with permission from Crystal Growth & Design, 2014, 14, 4562–4573. Copyright 2017, American Chemical Society.
Figure 19. Hydrogen bond interaction present in dapsone-caffeine 1:1 cocrystal. Adapted with permission from Crystal Growth & Design, 2014, 14, 4562–4573. Copyright 2017, American Chemical Society.
Crystals 08 00101 g019
Figure 20. (a) A zipper like structure in EA:GA cocrystal Form I. (b) 3D crystal packing showing the EA molecule connecting the zipper-like network. Adapted with permission from CrystEngComm, 2009, 11, 1823–1827. Copyright 2009, The Royal Society of Chemistry.
Figure 20. (a) A zipper like structure in EA:GA cocrystal Form I. (b) 3D crystal packing showing the EA molecule connecting the zipper-like network. Adapted with permission from CrystEngComm, 2009, 11, 1823–1827. Copyright 2009, The Royal Society of Chemistry.
Crystals 08 00101 g020
Figure 21. (a) A hydrogen bonded linear tape in form II crystal structure. (b) A hydrogen bonded tetrameric motif in EA-GA form III crystal structure. Adapted with permission from CrystEngComm, 2009, 11, 1823–1827. Copyright 2009, The Royal Society of Chemistry.
Figure 21. (a) A hydrogen bonded linear tape in form II crystal structure. (b) A hydrogen bonded tetrameric motif in EA-GA form III crystal structure. Adapted with permission from CrystEngComm, 2009, 11, 1823–1827. Copyright 2009, The Royal Society of Chemistry.
Crystals 08 00101 g021
Figure 22. Dissolution profile of EA-GA cocrystal polymorphs I, II, and III compared to pure drug EA. Adapted with permission from CrystEngComm, 2009, 11, 1823–1827. Copyright 2009, The Royal Society of Chemistry.
Figure 22. Dissolution profile of EA-GA cocrystal polymorphs I, II, and III compared to pure drug EA. Adapted with permission from CrystEngComm, 2009, 11, 1823–1827. Copyright 2009, The Royal Society of Chemistry.
Crystals 08 00101 g022
Figure 23. Acid-acid and acid-amide synthon observed in (a) FFA-CNB and (b) FFA-EA 1:1 cocrystals.
Figure 23. Acid-acid and acid-amide synthon observed in (a) FFA-CNB and (b) FFA-EA 1:1 cocrystals.
Crystals 08 00101 g023
Figure 24. (a) Hydrogen bond interaction present in the 1:1 FFA-theophylline cocrystal. (b) Intrinsic dissolution rate of FFA and theophylline cocrystals with theophylline and nicotinamide respectively. Adapted with permission from CrystEngComm, 2014, 16, 5793–5801. Copyright 2014, The Royal Society of Chemistry.
Figure 24. (a) Hydrogen bond interaction present in the 1:1 FFA-theophylline cocrystal. (b) Intrinsic dissolution rate of FFA and theophylline cocrystals with theophylline and nicotinamide respectively. Adapted with permission from CrystEngComm, 2014, 16, 5793–5801. Copyright 2014, The Royal Society of Chemistry.
Crystals 08 00101 g024
Figure 25. Hydrogen bond synthon present in (a) diclofenac–theophylline and (b) diflunisal-theophylline cocrystal. Adapted with permission from Mol. Pharmaceutics, 2014, 11 (10), 3707–3715. Copyright 2014, American Chemical Society.
Figure 25. Hydrogen bond synthon present in (a) diclofenac–theophylline and (b) diflunisal-theophylline cocrystal. Adapted with permission from Mol. Pharmaceutics, 2014, 11 (10), 3707–3715. Copyright 2014, American Chemical Society.
Crystals 08 00101 g025
Figure 26. (a) O–H···N hydrogen bond between COOH group of furosemide and imidazole N of caffeine. (b) Infinite chain of N–H···O hydrogen bond connecting C=O group of caffeine and primary sulfonamide NH. (c) Secondary NH bonded to sulfonyl O of another furosemide molecule. Adapted with permission from Journal of Pharmaceutical Sciences, 2012, 101, 664–680. Copyright 2012, John Wiley & Sons.
Figure 26. (a) O–H···N hydrogen bond between COOH group of furosemide and imidazole N of caffeine. (b) Infinite chain of N–H···O hydrogen bond connecting C=O group of caffeine and primary sulfonamide NH. (c) Secondary NH bonded to sulfonyl O of another furosemide molecule. Adapted with permission from Journal of Pharmaceutical Sciences, 2012, 101, 664–680. Copyright 2012, John Wiley & Sons.
Crystals 08 00101 g026
Figure 27. (a) Association of water molecule with two molecules of gefitinib through N–H···O, O–H···N and C–H···O interactions and a molecule of furosemide via O–H···O interaction. (b) API-API dimer between gefitinib and furosemide connecting carboxylic acid and morpholine group. Adapted with permission from Journal of Pharmaceutical Sciences, 2015, 104, 4207–4216. Copyright 2015, John Wiley & Sons.
Figure 27. (a) Association of water molecule with two molecules of gefitinib through N–H···O, O–H···N and C–H···O interactions and a molecule of furosemide via O–H···O interaction. (b) API-API dimer between gefitinib and furosemide connecting carboxylic acid and morpholine group. Adapted with permission from Journal of Pharmaceutical Sciences, 2015, 104, 4207–4216. Copyright 2015, John Wiley & Sons.
Crystals 08 00101 g027
Figure 28. Asymmetric unit containing MET (right) and GLI (left) molecules drawn at 50% probability level. Adapted with permission from Crystal Growth & Design, 2016, 16, 3577–3581. Copyright 2016, American Chemical Society.
Figure 28. Asymmetric unit containing MET (right) and GLI (left) molecules drawn at 50% probability level. Adapted with permission from Crystal Growth & Design, 2016, 16, 3577–3581. Copyright 2016, American Chemical Society.
Crystals 08 00101 g028
Figure 29. Infinite 1D hydrogen bond chain of MET-GLI crystal along the a-axis (blue for MET and red for GLI). Detailed interaction between MET-MET (right) and MET-GLI (left) presented in default color according to the different elements. Blue-dashed lines indicate hydrogen bonds. Adapted with permission from Crystal Growth & Design, 2016, 16, 3577–3581. Copyright 2016, American Chemical Society.
Figure 29. Infinite 1D hydrogen bond chain of MET-GLI crystal along the a-axis (blue for MET and red for GLI). Detailed interaction between MET-MET (right) and MET-GLI (left) presented in default color according to the different elements. Blue-dashed lines indicate hydrogen bonds. Adapted with permission from Crystal Growth & Design, 2016, 16, 3577–3581. Copyright 2016, American Chemical Society.
Crystals 08 00101 g029
Figure 30. DVS of (a) MET, (b) GLI, and (c) GLI-MET salt showing hydration stability (left) and (d) higher dissolution rate of GLI-MET salt compared to pure drug GLI (right). Adapted with permission from Crystal Growth & Design, 2016, 16, 3577–3581. Copyright 2016, American Chemical Society.
Figure 30. DVS of (a) MET, (b) GLI, and (c) GLI-MET salt showing hydration stability (left) and (d) higher dissolution rate of GLI-MET salt compared to pure drug GLI (right). Adapted with permission from Crystal Growth & Design, 2016, 16, 3577–3581. Copyright 2016, American Chemical Society.
Crystals 08 00101 g030
Figure 31. (a) Homodimer synthons of HCT and INZ. (b) 3D supramolecular structure of HCT-INZ-H2O. Adapted with permission from Crystal Growth & Design, 2017, 17, 308–316. Copyright 2017, American Chemical Society.
Figure 31. (a) Homodimer synthons of HCT and INZ. (b) 3D supramolecular structure of HCT-INZ-H2O. Adapted with permission from Crystal Growth & Design, 2017, 17, 308–316. Copyright 2017, American Chemical Society.
Crystals 08 00101 g031
Figure 32. 1D ribbon-type arrangement in lamivudine–zidovudine hydrate cocrystal. Adapted with permission from Crystal Growth & Design, 2009, 9, 951–957. Copyright 2009, American Chemical Society.
Figure 32. 1D ribbon-type arrangement in lamivudine–zidovudine hydrate cocrystal. Adapted with permission from Crystal Growth & Design, 2009, 9, 951–957. Copyright 2009, American Chemical Society.
Crystals 08 00101 g032
Figure 33. Schematic diagram showing the two drug molecules myricetin and piracetam, resulting in a 1:1 cocrystal structure using both liquid assisted grinding and solution crystallization. Adapted with permission from Journal of Molecular Structure, 2014, 1058, 114–121. Copyright 2014, Elsevier.
Figure 33. Schematic diagram showing the two drug molecules myricetin and piracetam, resulting in a 1:1 cocrystal structure using both liquid assisted grinding and solution crystallization. Adapted with permission from Journal of Molecular Structure, 2014, 1058, 114–121. Copyright 2014, Elsevier.
Crystals 08 00101 g033
Figure 34. (a) Hydrogen bond interaction present in NF-ST salt hydrate; (b) packing diagram showing water channel present in the salt hydrate. Adapted with permission from Molecular Pharmaceutics, 2016, 13, 3590–3594. Copyright 2016, American Chemical Society.
Figure 34. (a) Hydrogen bond interaction present in NF-ST salt hydrate; (b) packing diagram showing water channel present in the salt hydrate. Adapted with permission from Molecular Pharmaceutics, 2016, 13, 3590–3594. Copyright 2016, American Chemical Society.
Crystals 08 00101 g034
Figure 35. Enhanced solubility and dissolution rate of NF-ST salt hydrate compared to individual drug molecules. Adapted with permission from Molecular Pharmaceutics, 2016, 13, 3590–3594. Copyright 2016, American Chemical Society.
Figure 35. Enhanced solubility and dissolution rate of NF-ST salt hydrate compared to individual drug molecules. Adapted with permission from Molecular Pharmaceutics, 2016, 13, 3590–3594. Copyright 2016, American Chemical Society.
Crystals 08 00101 g035
Figure 36. Intensified method for cocrystal preparation. Adapted with permission from Cryst. Growth Des., DOI: 10.1021/acs.cgd.7b01197. Copyright 2018, American Chemical Society.
Figure 36. Intensified method for cocrystal preparation. Adapted with permission from Cryst. Growth Des., DOI: 10.1021/acs.cgd.7b01197. Copyright 2018, American Chemical Society.
Crystals 08 00101 g036
Figure 37. A four-component supramolecular unit in the crystal structure of OXP-SAL+-H salt. Adapted with permission from RSC Advances, 2016, 6, 34110–34119. Copyright 2016, The Royal Society of Chemistry.
Figure 37. A four-component supramolecular unit in the crystal structure of OXP-SAL+-H salt. Adapted with permission from RSC Advances, 2016, 6, 34110–34119. Copyright 2016, The Royal Society of Chemistry.
Crystals 08 00101 g037
Figure 38. Tablet resulting from compression experiments involving paracetamol (pca form I) and cocrystals with theophylline (thp), naphthalene (nap), oxalic acid (oxa), and phenazine (phe) respectively. Adapted with permission from Adv. Mater. 2009, 21, 3905–3909. Copyright 2009, John Wiley & Sons.
Figure 38. Tablet resulting from compression experiments involving paracetamol (pca form I) and cocrystals with theophylline (thp), naphthalene (nap), oxalic acid (oxa), and phenazine (phe) respectively. Adapted with permission from Adv. Mater. 2009, 21, 3905–3909. Copyright 2009, John Wiley & Sons.
Crystals 08 00101 g038
Figure 39. Probable hydrogen bond synthon (a) acid-amide and (b) acid-pyridine possible between the two drug molecules in 1:1 cocrystal. Adapted with permission from Crystal Growth & Design, 2011, 11, 4780–4788. Copyright 2011, American Chemical Society.
Figure 39. Probable hydrogen bond synthon (a) acid-amide and (b) acid-pyridine possible between the two drug molecules in 1:1 cocrystal. Adapted with permission from Crystal Growth & Design, 2011, 11, 4780–4788. Copyright 2011, American Chemical Society.
Crystals 08 00101 g039
Scheme 3. Molecular structure of pyrimethamine and coformers used in the study with outcome of cocrystallization studies. Adapted with permission from CrystEngComm, 2013, 15, 2916–2928. Copyright 2013, The Royal Society of Chemistry.
Scheme 3. Molecular structure of pyrimethamine and coformers used in the study with outcome of cocrystallization studies. Adapted with permission from CrystEngComm, 2013, 15, 2916–2928. Copyright 2013, The Royal Society of Chemistry.
Crystals 08 00101 sch003
Figure 40. Hydrogen bond interaction present in 1:1 temozolomide cocrystal of (a) pyrazinamide and (b) caffeine. Adapted with permission from Crystal Growth & Design, 2013, 13 (5), 2208–2219. Copyright 2013, American Chemical Society.
Figure 40. Hydrogen bond interaction present in 1:1 temozolomide cocrystal of (a) pyrazinamide and (b) caffeine. Adapted with permission from Crystal Growth & Design, 2013, 13 (5), 2208–2219. Copyright 2013, American Chemical Society.
Crystals 08 00101 g040
Figure 41. (a) Hydrogen bond interaction present in theophylline–acetazolamide cocrystal structure. (b) Plastic deformation in theophylline-acetazolamide cocrystal before (left) and after (right) application of mechanical stress. Adapted with permission from CrystEngComm, DOI: 10.1039/C7CE01915G. Copyright 2018, The Royal Society of Chemistry.
Figure 41. (a) Hydrogen bond interaction present in theophylline–acetazolamide cocrystal structure. (b) Plastic deformation in theophylline-acetazolamide cocrystal before (left) and after (right) application of mechanical stress. Adapted with permission from CrystEngComm, DOI: 10.1039/C7CE01915G. Copyright 2018, The Royal Society of Chemistry.
Crystals 08 00101 g041
Figure 42. Design and synthesis of VD2-VD3 cocrystal. Blue and green ball represent the substituent groups present in VD2 and VD3. Adapted with permission from Chemical Communications, 2016, 52, 3572–3575. Copyright 2016, The Royal Society of Chemistry.
Figure 42. Design and synthesis of VD2-VD3 cocrystal. Blue and green ball represent the substituent groups present in VD2 and VD3. Adapted with permission from Chemical Communications, 2016, 52, 3572–3575. Copyright 2016, The Royal Society of Chemistry.
Crystals 08 00101 g042
Figure 43. Packing arrangement of cocrystal form A and form B viewed along the helix structure. Adapted with permission from Chemical Communications, 2016, 52, 3572–3575. Copyright 2016, The Royal Society of Chemistry.
Figure 43. Packing arrangement of cocrystal form A and form B viewed along the helix structure. Adapted with permission from Chemical Communications, 2016, 52, 3572–3575. Copyright 2016, The Royal Society of Chemistry.
Crystals 08 00101 g043
Figure 44. Change in vitamin D3 assay during storage at (a) 23 °C with an illuminance of 5000 lx and (b) 40 °C/75%RH for cocrystal A and B compared with starting materials. Adapted with permission from Chemical Communications, 2016, 52, 3572–3575. Copyright 2016, The Royal Society of Chemistry.
Figure 44. Change in vitamin D3 assay during storage at (a) 23 °C with an illuminance of 5000 lx and (b) 40 °C/75%RH for cocrystal A and B compared with starting materials. Adapted with permission from Chemical Communications, 2016, 52, 3572–3575. Copyright 2016, The Royal Society of Chemistry.
Crystals 08 00101 g044
Figure 45. Comparison between the aqueous solubility of pterostilbene and the solution concentration of pterostilbene achieved from three cocrystals with carbamazepine, piperazine and caffeine. Adapted with permission from Crystal Growth & Design, 2011, 11, 2817–2823. Copyright 2011, American Chemical Society.
Figure 45. Comparison between the aqueous solubility of pterostilbene and the solution concentration of pterostilbene achieved from three cocrystals with carbamazepine, piperazine and caffeine. Adapted with permission from Crystal Growth & Design, 2011, 11, 2817–2823. Copyright 2011, American Chemical Society.
Crystals 08 00101 g045
Scheme 4. Molecular structure of a few nutraceutical molecules widely used as cocrystal former.
Scheme 4. Molecular structure of a few nutraceutical molecules widely used as cocrystal former.
Crystals 08 00101 sch004
Figure 46. Various cocrystals of isoniazid prepared using nutraceuticals as coformer. Adapted with permission from Crystal Growth & Design, 2014, 14, 5991–6005. Copyright 2014, American Chemical Society.
Figure 46. Various cocrystals of isoniazid prepared using nutraceuticals as coformer. Adapted with permission from Crystal Growth & Design, 2014, 14, 5991–6005. Copyright 2014, American Chemical Society.
Crystals 08 00101 g046
Figure 47. Bar diagram showing solubility profile of (a) ethenzamide cocrystals and (b) Pyz binary complexes. Ferulic acid is designated as FRA. Adapted with permission from CrystEngComm, 2017, 19, 826–833. Copyright 2017, The Royal Society of Chemistry.
Figure 47. Bar diagram showing solubility profile of (a) ethenzamide cocrystals and (b) Pyz binary complexes. Ferulic acid is designated as FRA. Adapted with permission from CrystEngComm, 2017, 19, 826–833. Copyright 2017, The Royal Society of Chemistry.
Crystals 08 00101 g047
Figure 48. Dissolution profile of quercetin dihydrate and its cocrystals in 1:1 ethanol/water mixture after the first 4 h. Adapted with permission from Mol. Pharmaceutics, 2011, 8 (5), 1867–1876. Copyright 2011, American Chemical Society.
Figure 48. Dissolution profile of quercetin dihydrate and its cocrystals in 1:1 ethanol/water mixture after the first 4 h. Adapted with permission from Mol. Pharmaceutics, 2011, 8 (5), 1867–1876. Copyright 2011, American Chemical Society.
Crystals 08 00101 g048
Table 1. Results of various solution crystallization experiments
Table 1. Results of various solution crystallization experiments
PolymorphSolvent of Crystallization
Form IAcetonitrile
Formic acid
Toluene + ethylacetate
Form II2-Propanol
Toluene + acetonitrile
Chloroform + ethylacetate
Ethylacetate + n-hexane
Ethylacetate + cyclohexane
Form IIIAcetone
Acetonitrile + acetone
Methanol + chloroform
Chloroform + acetonitrile
Toluene + acetonitrile

Share and Cite

MDPI and ACS Style

Thakuria, R.; Sarma, B. Drug‑Drug and Drug‑Nutraceutical Cocrystal/Salt as Alternative Medicine for Combination Therapy: A Crystal Engineering Approach. Crystals 2018, 8, 101. https://doi.org/10.3390/cryst8020101

AMA Style

Thakuria R, Sarma B. Drug‑Drug and Drug‑Nutraceutical Cocrystal/Salt as Alternative Medicine for Combination Therapy: A Crystal Engineering Approach. Crystals. 2018; 8(2):101. https://doi.org/10.3390/cryst8020101

Chicago/Turabian Style

Thakuria, Ranjit, and Bipul Sarma. 2018. "Drug‑Drug and Drug‑Nutraceutical Cocrystal/Salt as Alternative Medicine for Combination Therapy: A Crystal Engineering Approach" Crystals 8, no. 2: 101. https://doi.org/10.3390/cryst8020101

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