Next Article in Journal
l-Lysine-Based Gelators for the Formation of Gels in Water and Alcohol–Water Mixtures
Next Article in Special Issue
Cryopreservation of 3D Bioprinted Scaffolds with Temperature-Controlled-Cryoprinting
Previous Article in Journal
Growth Factor Loaded Thermo-Responsive Injectable Hydrogel for Enhancing Diabetic Wound Healing
Previous Article in Special Issue
A Power Compensation Strategy for Achieving Homogeneous Microstructures for 4D Printing Shape-Adaptive PNIPAM Hydrogels: Out-of-Plane Variations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Thermally Resistant Bisamide Gelators as Pharmaceutical Crystallization Media

1
Department of Organic Chemistry, Faculty of Chemistry, University of Murcia, 30100 Murcia, Spain
2
Department of Organic Chemistry, Faculty of Chemical Sciences and Technologies, University of Castilla La Mancha-IRICA, 13071 Ciudad Real, Spain
3
Department of Chemistry, Durham University, South Road, Durham DH1 3LE, UK
4
Department of Chemical Sciences, Tezpur University, Napaam 784028, India
*
Authors to whom correspondence should be addressed.
Submission received: 31 October 2022 / Revised: 12 December 2022 / Accepted: 27 December 2022 / Published: 29 December 2022
(This article belongs to the Special Issue 3D Printing of Gels: Applications and Properties)

Abstract

:
Three simple bisamide derivatives (G1, G2 and G3) with different structural modifications were synthesized with easy synthetic procedures in order to test their gel behaviour. The outcomes showed that hydrogen bonding was essential in gel formation; for this reason, only G1 provided satisfactory gels. The presence of methoxy groups in G2 and the alkyl chains in G3 hindered the hydrogen bonding between N-H and C=O that occurred G1. In addition, G1 provided thermally and mechanical stable gels, as confirmed with Tsol and rheology experiments. The gels of G1 were also responsive under pH stimuli and were employed as a vehicle for drug crystallization, causing a change in polymorphism in the presence of flufenamic acid and therefore providing the most thermodynamically stable form III compared with metastable form IV obtained from solution crystallization.

Graphical Abstract

1. Introduction

The booming interest in supramolecular gels in the last decade contrasts with their initially slow development following their original discovery in the 1930s [1]. In particular, low-molecular-weight gelators (LMWGs) have aroused great interest and comprise organic or coordination compounds with a molecular weight lower than 2000 Da; LMWGs are capable of gelling organic solvents or water. The interest in LMWGs arises from their versatility, good solubility upon heating, and ability to gel at low gelator concentrations (often less than 1% by mass), as well as the gels’ thermally reversible sol-to-gel phase transition. Gels obtained from LMWGs are usually prepared by cooling a heated solution of LMWGs in the solvent to be gelled, leading to a supersaturated solution. This supersaturation induces a rapid assembly of the gelator molecules into elongated fibres of typically 5–100 nm in diameter; these fibres subsequently aggregate into a 3D entangled network, thereby turning the liquid into a supramolecular gel [2]. Hence, supramolecular gels consist of an entangled fibrous network of gelator molecules held together by specific intermolecular interactions between gelators. Despite the fact that there are many different LMWGs with very different structures, it is still a challenge to predict the molecular structure of a possible gelator and the solvents in which it might form a gel. In general, the self-assembly of gelators into fibrous networks is driven by non-covalent interactions, such as van der Waals interactions, ππ interactions, dipolar interactions, hydrogen bonding, or coulombic interactions. Furthermore, solvophobic effects also play an important role because they contribute to the gelating ability by reducing the overall solubility of a gelator in a specific solvent [3,4,5].
To be able to design new gelators or even to predict gelation behaviour from molecular structure requires basic insight into the physicochemical basis for their gelation behaviour [6]. The problem is that these systems and the structures of the gels themselves have a large structural diversity. Many different techniques have been applied to establish the intramolecular interactions involved in self-assembly leading to gelation in order to elucidate the structures and morphologies of gels and fibres and to determine their thermotropic and viscoelastic properties [7].
One emerging application of low-molecular-weight organogels in the last decade is as media for the crystallization of pharmaceuticals. The control of the solid form of a drug and its properties, particularly bioavailability, is crucial in the pharmaceutical industry [8,9,10]. The crystallization of the active ingredient is of vital importance because the polymorphic form, the crystal size, or the crystal morphology can influence its solubility, compressibility, melting point, bulk density, and dissolution rate [11,12]. In this sense, gels can be interesting media for polymorph discovery [13,14,15] and may offer interesting benefits such as the control of a specific polymorphic form, thus offering a potential alternative nucleation pathway compared with conventional solution crystallization. As a result, there is growing interest in organic compounds that have the ability to form gels and that can interact with a target pharmaceutical compound in such a way as to influence its crystallization outcome. However, the preparation of a gelator with appropriate functionalities can involve a complicated multi-step synthetic procedure. In this work, we describe the preparation of three simple bisamide-derived LMWGs (G1–3) (Scheme 1) and their application in pharmaceutical crystallization. It is anticipated that the presence of amide or methoxy groups may facilitate the formation of gels through hydrogen bonding [16]. Likewise, the presence of the alkyl chains in G3 confers flexibility and hydrophobicity to the molecule and facilitates the gelation process [17].

2. Results and Discussion

2.1. Synthetic Procedure

Compound G1, which was previously described by Shanmuga et al. for different purposes [18], was synthesized for the purposes of this work, and G2 were prepared from acid chloride and anhydrous hydrazine [19]. Compound G3 was prepared from G1 through an alkylation reaction employing KOH and hexyl bromide (Scheme 1) [20]. All compounds provided satisfactory characterization data.

2.2. Gelation Tests

Gelation tests for compounds G13 were performed with thirty different solvents and at different concentrations (2%, 1% and 0.5% wt.). Solvents were chosen to span the polarity spectrum from water to hexane and included aliphatic media; electron-rich and electron-poor aromatic solvents; and protic, dipolar aprotic, and halogenated systems. The gelators were dissolved in 0.5 mL of the chosen solvent followed by sonication for 30–60 s until complete dissolution and controlled heating at a temperature slightly below the boiling point of the solvent for a minute. Then, the vials were left undisturbed at room temperature and were tested for gelation at various time intervals: 30 min, 4 h, 24 h, 48 h and 72 h. Simple tube inversion was used to detect gel formation. The results are summarized in Supporting Information in Tables S1–S9.
The results of the gelation tests showed that compound G1 provided a range of gels, while G2 and G3 did not gel in any solvent. Gelator G1 formed gels in nine different solvents (dichloromethane, ethanol, 1-pentanol, 1-propanol, 2-butanol, acetone, acetonitrile, water and ethyl acetate) at a concentration of 2% (Figure 1a) and in five solvents (dichloromethane, 1-pentanol, 1-propanol, water and ethyl acetate) at 1% wt. (Figure 1b). It remains very challenging to produce and rationalize gelation behaviour in particular solvents, and the solvents that formed gels here were those in which the compound had intermediate solubility that allowed for fibre formation when hot but resulted in insolubility when cold. This compound proved to be highly effective in the gelation processes, especially in polar solvents such as water, probably due to the presence of the bisamide group that enabled the formation of hydrogen bonding. In contrast, the presence of the aliphatic chains in G3 avoided the formation of hydrogen bonds. G2 showed a curious behaviour, since it did not form any gels despite having amide and methoxyl groups in its structure that could facilitate the formation of hydrogen bonds.
In order to probe the aggregation of the candidate gelators, the structures of G1 and G2 were determined with single-crystal X-ray diffraction. The crystal structure of G1 was previously described by Shanmuga et al. [18]. Unfortunately, no suitable single-crystals of G3 were obtained. In G1, the main interactions in the solid state structure were found to be hydrogen bonds between the N-H and C=O groups (Figure 2). Compound G2 crystallized as a dihydrate from ethanol. The steric bulk of the trimethoxyaryl groups limited the ability of compound G2 to form a hydrogen-bonded chain between the N-H and C=O groups, and the NH···O interactions were longer than in G1 at 2.99 Å (Figure 3). The water of hydration formed an infinite chain linking one amide chain to another. The longer and hence weaker hydrogen-bonded chain interactions and the need to include water in the structure of G2 may offer an explanation for its lack of gelation behaviour.
Critical gelation concentration (CGC) is an essential parameter for determining the capacity of a gelator to form a gel [21]. It can be defined as the minimum concentration of a gelator needed to form a consistent gel. These values are provided in Table 1.
The CGC for G1 ranged between 0.6 in the cases of dichloromethane and ethyl acetate and 1.5 in the cases of ethanol and acetonitrile. The CGCs were found to be relatively low for very polar solvents such as ethyl acetate and dichloromethane or more hydrophobic alcohols such as 1-pentanol and 1-propanol.
The gel-to-sol phase transition temperature (Tsol) was also determined for the gels of G1 by employing the dropping ball method [22,23] and the results were recorded in Table 2. In this method, a small ball is deposited in the middle of a gel surface. Then, the temperature is slowly increased until the gel transforms into a sol and the ball touches the ground of the vial. The temperature at this point is Tsol. The ball must be inert with respect to the gelator and not too heavy or too light to avoid the ball sinking into the gel or floating on the solution surface. Some gels of G1 were found to be very stable, for example, in 1-pentanol, 2-butanol, and (especially) water or ethyl acetate, with values above the solvent boiling point in some cases. The strong hydrogen bonds might explain the thermal strength of these gels and the high values of Tsol. Reductions in the concentration led to decreases in Tsol as expected, but in some solvents such as dichloromethane and 1-propanol, the stability was practically the same regardless of the concentration.
The morphology of the supramolecular aggregates in the form of dried xerogels were studied with SEM [24,25]. As usual, gels from G1 showed an interconnected fibrous system [26] (Figure 4a). On the other hand, an amorphous morphology was observed for G3, which was consistent with the lack of gelation of this compound (Figure 4b).
To study the resistance under mechanical stimuli of the gels, we performed rheology experiments of gels in different solvents in which gels were formed at the same concentration (2% wt.) and compared the stability of the different gels for the same solvent at different concentrations. Frequency sweep experiments were performed and showed that in all the gels, the elastic modulus, G′, was at least an order of magnitude greater than the viscous modulus, G″. The experiments also showed that G′ was practically unchanged regardless of frequency, thus confirming the gel behaviour observed in the previously described inversion tube tests. The parameter that was used for the comparison was the yield stress (σ), which can be defined as the point which gels break down under mechanical stimuli and begin to shear. As a general rule, the stability of gels is higher with high values of yield stress [27].
The gels of G1 proved to be highly robust, with σ values of more than 1000 Pa in 1-pentanol and ethanol, as shown in Figure 5. The low σ values in the case of water are remarkable, and the mechanical stability of the hydrogel was relatively low. The results of the remaining stress sweep experiments are described in the SI (Figures S1–S7).
In addition, gels in the same solvent (1-pentanol) were compared at different concentrations (2% vs. 1% wt.) (Figure 5b vs. Figure S8 in the SI). In the same way as the Tsol experiments, the mechanical stability of the gels (confirmed by the decreasing σ) decreased with the concentration because the number of interconnections between the fibres decreased due to the lower fibre density causing the weakening of the gel.

2.3. Stimuli Response

Gel responsiveness under different stimuli is of interest in the context of smart materials [28,29]. The importance of hydrogen-bonding interactions in the organogels of G1 led us to study the influence of pH on its gelation process. G1 gels in ethanol at a concentration of 2% wt. were treated with a solution of NaOH 0.1 M (pH = 13) to induce the transformation of the gel into a sol (Figure 6). The results of this treatment indicated the breaking of the hydrogen bond network due to the deprotonation of the bisamide group or hydrogen bond competition by the excess hydroxide ions. This hypothesis was formulated after the treatment of this mixture with a solution of HCl 0.1M (pH = 1). In this case, the gel recovered its initial structure (Figure 6), again indicating re-protonation and confirming the great importance of hydrogen bonds in the formation of G1 gels.

2.4. Gel Phase Crystallization of Pharmaceutical Drugs

Gels with efficient thermal and mechanical resistance can be applied in different applications as drug delivery or drug crystallization systems. Hydrogels are suitable for drug delivery due to their potential biocompatibility caused by the presence of water in their structures. Despite the fact that G1 can form hydrogels, its low mechanical resistance confirmed by rheological studies indicates that this gel is not suitable as a drug delivery vehicle. However, gels of G1 are potentially suitable for the crystallization of active pharmaceutical ingredients (API) [8,9,10]. Small-molecule organogels reduce convection (thus allowing for diffusional growth) and have the possibility of functional-group-specific interactions with the gel surface, potentially allowing the gel to act as a heteronucleation surface. Based on the structure of G1, we screened the crystallization of potentially complementary active pharmaceutical ingredients (APIs) in gels. We chose APIs with efficient NH-type hydrogen bonding functional groups, known polymorphism, and aromatic groups to potentially π-stack with the gel surface, namely, theophylline, sulfamerazine, sulfathiazole, and flufenamic acid (Figure 7).
A wide range of different crystallization tests were carried out to optimize the crystallization conditions of the APIs. The best gelling conditions were found when using a 10 mg/mL drug concentration at a 1% G1 concentration. In a vial, a mixture of G1 and the different APIs in a specific solvent was heated, sonicated, and kept undisturbed at room temperature for up to several weeks to allow crystallization to complete. In many cases, crystallization was not observed. The tests were repeated in the same conditions five times. The vials were visually checked for the presence of solid materials, and the resulting crystals were analysed with X-ray powder and single-crystal X-ray crystal diffraction. The G1 gelator showed the ability to act as a successful host for the crystallization of all APIs; however, the same polymorphic form was observed in both a control solution and the gel phase crystallization for sulfathiazole, theophylline, and sulfamerazine. The outcomes showed the existence of kinetic form II in the case of theophylline in dichloromethane [30,31], the Pna21 polymorph in the case of sulfamerazine [32], and polymorph II in the case of sulfathiazole [33,34]. On the other hand, in the gel crystallization of flufenamic acid, a change in the polymorphism from form IV in solution to form III was observed in the presence of the G1 gelator [35]. The polymorphism was confirmed with the unit cell determination of the crystals (data in the SI). This change in the polymorphism was also accompanied by a change in the morphology of the crystals from block crystals in solution evaporation to thick needles crystals in gel crystallization (Figure 8) These results suggested that the presence of G1 inhibited the crystallization of form IV and favoured the nucleation of form III of flufenamic acid, perhaps as a result of the common acid–amide heterosynthon [36]. Form III is the most thermodynamically stable form at room temperature, and it is created with stirred cooling of a hot toluene solution [37]. Its observation under conditions that favour a metastable polymorph suggests that the studied gels may be useful in sidestepping the formation of metastable forms to target the pharmaceutically desirable thermodynamic phase.

3. Conclusions

Three readily synthesized bisamide derivatives were tested as possible gelators. Compound G1 proved to be an effective gelator with a wide range of gels in different solvents at different concentrations with response under pH stimuli. Some gels of G1 exhibited high thermal and mechanical resistance. In addition, gels of G1 were employed in the crystallization of theophylline, sulfamerazine, sulfathiazole, and flufenamic acid. In this last case, the presence of the gel induced a change in the polymorphism to the most thermodynamically stable form, form III, under ambient conditions compared with a control solution under the same conditions. These results suggest the potential of gels to ‘sidestep’ metastable forms and allow for the direct crystallization of the pharmaceutically desirable thermodynamic form.

4. Materials and Methods

All reagents were used without any previous purification. Reactions with materials sensitive to air were performed under an inert atmosphere of argon. Flash chromatography was performed with silica gel (Merck, Kieselgel 60, 230–240 mesh or Scharlau 60, 230–240 mesh). Analytical thin-layer chromatography (TLC) was carried out with the use of aluminium-coated Merck Kieselgel 60 F254 plates. NMR spectra were recorded with a Varian Unity 500 (1H: 500 MHz; 13C: 125 MHz) spectrometer at 298 K while employing deuterated solvents as internal references against the residual protic solvent signal. Chemical shifts (δ) are described in ppm. Multiplicities are denoted in the following way: s = singlet; d = doublet; t = triplet; m = multiplet; br = broad.
Melting points were determined with a Buchi Melting-point 565 instrument.
SEM images were obtained with a JEOL JSM 6335F microscope working at 10 kV.
In the drop-milling test used to calculate Tsols, a small metal ball with a diameter of 1 cm was used.
Rheological measurements were carried out with advanced AR 2000 rheometer from TA Instruments that was equipped with a cooling system (Julabo C). A 20 mm plain plate geometry (stainless steel) was also employed. Firstly, strain sweep measurements were carried out in order to estimate the strain in percent at which reasonable torque values were given (about 10 times of the transducer resolution limit). Then, frequency sweep measurements and time sweep measurements from 0.1 to 4000 Pa were carried out.
Gel phase crystallizations were carried out with the addition of 10 mg of the drug in a vial containing 1% wt. of the G1 gelator in a particular solvent. The vials were warmed and then sonicated to achieve a gel of the API solution. Crystals started appearing after a few days, as observed with microscopy. Crystals were characterized with the determination of the unit cell parameter.
The single-crystal X-ray data for G2 dihydrate were collected at a temperature of 120.0(2)K using MoKα radiation (λ = 0.71073 Å) with a Bruker D8Venture (Photon100 CMOS detector, IμS-microsource, focusing mirrors) 3-circle diffractometer equipped with a Cryostream (Oxford Cryosystems) open-flow nitrogen cryostat. The structure was solved with the direct method and refined with full-matrix least squares on F2 for all data using Olex2 [38] and SHELXTL [39] software. All non-hydrogen atoms were refined in anisotropic approximation, and hydrogen atoms were found with difference Fourier map and isotropically refined. Crystal data and the parameters of refinement are summarized in the Supporting Information in Tables S10–S14. Crystallographic data for the structure were deposited in the Cambridge Crystallographic Data Centre as supplementary publication CCDC-2216521.

Synthetic Procedures for the Synthesis of G13

N’-benzoylbenzohydrazide (G1): In a two-necked flask, THF (20 mL), hydrazine (1.14 g, 35.71 mmol) and Et3N (20 mL) were added at 0 °C and under an inert atmosphere. Then, acid chloride (10.0 g, 71.42 mmol) dissolved in THF was added dropwise. The reaction mixture was stirred at room temperature overnight. A saturated solution of K2CO3 was added, and the mixture was stirred for 2 h. Finally, the reaction mixture was filtered to obtain a white solid (6.86 g, 80%) corresponding to G1 that was washed with water (3 × 20 mL). M.p: 238–240 °C. 1H-NMR (CDCl3, ppm) δ: 8.18 (s, 2H, NH), 8.03–8.01 (d, J = 7.4 Hz, 4H, o-Ph), 7.69–7.63 (m, 6H, m,p-Ph). 13C-NMR (CDCl3, ppm) δ: 164.9, 132.2, 131.9, 128.8, 127.7. The MS calculated for (C14H12N2O2) M+ 240.09 was found to be 240.98.
3,4,5-trimethoxy-N’-(3,4,5-trimethoxybenzoyl)benzohydrazide (G2): In a two-necked flask, THF (20 mL), hydrazine (0.70 g, 21.74 mmol) and Et3N (20 mL) were added at 0 °C under an inert atmosphere. Then, acid chloride (10.0 g, 43.47 mmol) dissolved in THF was added dropwise. The reaction mixture was stirred at room temperature overnight. A saturated K2CO3 solution was added, and the mixture was stirred for 2 h. Finally, the reaction mixture was filtered to obtain a white solid (6.85 g, 75%) corresponding to G2 that was washed with water (3 × 20 mL). M.p: 245–247 °C. 1H-NMR (CDCl3, ppm) δ: 8.16 (s, 2H, NH), 7.17 (s, 4H, o-Ph), 3.91 (s, 18H, -OCH3) 13C-NMR (CDCl3, ppm) δ: 164.8, 153.1, 142.7, 128.6, 126.6, 60.6, 56.4. The MS calculated for (C20H24N2O8) M+ 420.15 was found to be 420.65.
N’-benzoyl-N,N’-dihexylbenzohydrazide (G3): A mixture of G1 (5 g, 20.82 mmol), potassium tert-butoxide (2.71 g, 24.15 mmol), and 1-bromohexane (3.61 g, 21.86 mmol) was dissolved in methanol (50 mL). The reaction was heated at reflux for 12 h. The solvent was then removed under reduced pressure, and the residue was dissolved in CHCl3 and washed with H2O and brine. The organic phase was dried over MgSO4, the solvent was removed, and column chromatography was carried out while employing hexane/AcOEt 9:1 as the eluant, thus obtaining compound G3 as a pale yellow solid (6.21 g, 73%). M.p: 202–204°C. 1H-NMR (CDCl3, ppm) δ: 8-03-8.01 (d, J = 7.4 Hz,, 4H, o-Ph), 7.68–7.63 (m, 6H, m,p-Ph), 4.62 (t, J = 7.0 Hz, 4H, N-CH2), 1.56–1.52 (t, J = 7.0 Hz, 4H, -CH2), 1.31–1.26 (m, 12H, -CH2), 0.90 (t, J = 7.0 Hz, 6H, -CH3) 13C-NMR (CDCl3, ppm) δ: 171.5, 137.9, 132.1, 128.9, 127.5, 44.6, 28.7, 27.4, 27.2, 22.6, 14.6. The MS calculated for (C26H36N2O2) M+ 408.28 was found to be 408.05.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/gels9010026/s1. Tables S1–S9: Gelation test at different concentrations of G1G3; Figures S1–S8: Rheology experiments for G1; Tables S10–S17: Crystal data for G2. Reference [40] is cited in the supplementary materials

Author Contributions

Conceptualization, I.T.-M., J.R.C., P.P. and J.W.S.; methodology, I.T.-M., A.S. and B.S.; formal analysis and investigation, I.T.-M., A.S., D.S.Y. and B.S.; data curation, I.T.-M., D.S.Y. and B.S.; writing—original draft preparation, I.T.-M., J.R.C., P.P. and J.W.S.; writing—review and editing, I.T.-M. and J.W.S.; supervision, P.P. and J.R.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Ministerio de Ciencia e Innovación/Agencia Estatal de investigación” (MCIN/AEI) of Spain (project PID2020-106114GB-I00) and the Junta de Comunidades de Castilla La Mancha (JCCM-FEDER) (project SBPLY/21/180501/000114).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was financially supported by the Ministerio de Ciencia e Innovación/Agencia Estatal de investigación” (MCIN/AEI) of Spain (project PID2020-106114GB-I00) and the Junta de Comunidades de Castilla La Mancha (JCCM-FEDER) (project SBPLY/21/180501/000114). I. Torres-Moya is indebted to Juan de la Cierva Formación 2020 FJC2020-043684-I de la ayuda financiada por MCIN/AEI/10.13039/501100011033 y por la Unión Europea NextGeneration EU/PRTR. B. Saikia acknowledges the Commonwealth Scholarship Commission for a split-site PhD fellowship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jordan Lloyd, D. Colloid Chemistry. Alexander, J., Ed.; The Chemical Catalogue Company: New York, NY, USA, 1926; Volume 1, pp. 767–782. [Google Scholar]
  2. Terech, P.; Weiss, R.G. Low Molecular Mass Gelators of Organic Liquids and the Properties of Their Gels. Chem. Rev. 1997, 97, 3133–3159. [Google Scholar] [CrossRef] [PubMed]
  3. Draper, E.R.; Adams, D.J. Low-Molecular-Weight Gels: The State of the Art. Chem 2017, 3, 390–410. [Google Scholar] [CrossRef] [Green Version]
  4. Lan, Y.; Corradini, M.G.; Weiss, R.G.; Raghavan, S.R.; Rogers, M.A. To gel or not to gel: Correlating molecular gelation with solvent parameters. Chem. Soc. Rev. 2015, 44, 6035–6058. [Google Scholar] [CrossRef]
  5. Dastidar, P. Designing Supramolecular Gelators: Challenges, Frustrations, and Hopes. Gels 2019, 5, 15–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Hirst, A.R.; Coates, I.A.; Boucheteau, T.R.; Miravet, J.F.; Escuder, B.; Castelletto, V.; Hamley, I.W.; Smith, D.K. Low-molecular-weight gelators: Elucidating the principles of gelation based on gelator solubility and a cooperative self-assembly model. J. Am. Chem. Soc. 2008, 130, 9113–9121. [Google Scholar] [CrossRef]
  7. Yu, G.; Yan, X.; Han, C.; Huang, F. Characterization of supramolecular gels. Chem. Soc. Rev. 2013, 42, 6697–6722. [Google Scholar] [CrossRef]
  8. Foster, J.A.; Damodaran, K.K.; Maurin, A.; Day, G.M.; Thompson, H.P.G.; Cameron, G.J.; Bernal, J.C.; Steed, J.W. Pharmaceutical polymorph control in a drug-mimetic supramolecular ge. Chem. Sci. 2017, 8, 78–84. [Google Scholar] [CrossRef] [Green Version]
  9. Foster, J.A.; Piepenbrock, M.O.M.; Lloyd, G.O.; Clarke, N.; Howard, J.A.K.; Steed, J.W. Anion-switchable supramolecular gels for controlling pharmaceutical crystal growth. Nat. Chem. 2010, 2, 1037–1043. [Google Scholar] [CrossRef]
  10. Aparicio, F.; Matesanz, E.; Sanchez, L. Cooperative self-assembly of linear organogelators. Amplification of chirality and crystal growth of pharmaceutical ingredients. Chem. Commun. 2012, 48, 5757–5759. [Google Scholar] [CrossRef] [PubMed]
  11. Hilfiker, R. Polymorphism in the Pharmaceutical Industry; Wiley-VCH: Weinheim, Germany, 2006. [Google Scholar]
  12. Lee, E.H. Asian A practical guide to pharmaceutical polymorph screening & selection. J. Pharm. Sci. 2014, 9, 163–175. [Google Scholar] [CrossRef]
  13. Dawn, A.; Andrew, K.S.; Yufit, D.S.; Hong, Y.X.; Reddy, J.P.; Jones, C.D.; Aguilar, J.A.; Steed, J.W. Supramolecular Gel Control of Cisplatin Crystallization: Identification of a New Solvate Form Using a Cisplatin-Mimetic Gelator. Growth Des. 2015, 15, 4591–4599. [Google Scholar] [CrossRef] [Green Version]
  14. Buendía, J.; Matesanz, E.; Smith, D.K.; Sánchez, L. Multi-component supramolecular gels for the controlled crystallization of drugs: Synergistic and antagonistic effects. CrystEngComm 2015, 17, 8146–8152. [Google Scholar] [CrossRef] [Green Version]
  15. Kumar, D.K.; Steed, J.W. Supramolecular gel phase crystallization: Orthogonal self-assembly under non-equilibrium conditions. Chem. Soc. Rev. 2014, 43, 2080–2088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Rao, M.R.; Sun, S.-S. Supramolecular Assemblies of Amide-Derived Organogels Featuring Rigid π-Conjugated Phenylethynyl Frameworks. Langmuir 2013, 29, 15146–15158. [Google Scholar] [CrossRef] [PubMed]
  17. Moniruzzaman, M.; Sundarajan, P.R. Low Molecular Weight Organogels Based on Long-Chain Carbamates. Langmuir 2005, 21, 3802–3807. [Google Scholar] [CrossRef] [PubMed]
  18. Shanmuga Sundara Raj, S.; Yamin, B.M.; Boshaala, A.M.A.; Tarafder, M.T.H.; Crouse, K.A.; Fun, H.-K. 1,2-Dibenzoylhydrazine. Acta Cryst. 2000, 56, 1011–1012. [Google Scholar] [CrossRef] [Green Version]
  19. Kumar, M.M.; Venkataramana, P.; Swamy, P.Y.; Chityala, Y. N-Amino-1,8-Naphthalimide is a Regenerated Protecting Group for Selective Synthesis of Mono-N-Substituted Hydrazines and Hydrazides. Chem. Eur. J. 2021, 27, 17713–17721. [Google Scholar] [CrossRef]
  20. Ekiz, F.; Rende, E.; Timur, S.; Toppare, L. A novel functional conducting polymer: Synthesis and application to biomolecule immobilization. J. Mater. Chem. 2012, 22, 22517–22525. [Google Scholar] [CrossRef]
  21. Pekcan, O.; Kara, S. Gelation Mechanisms. Mod. Phys. Lett. B 2012, 26, 30019–30046. [Google Scholar] [CrossRef]
  22. Mukkamala, R.; Weiss, R.G. Physical Gelation of Organic Fluids by Anthraquinone−Steroid-Based Molecules. Structural Features Influencing the Properties of Gels. Langmuir 1996, 12, 1474–1482. [Google Scholar] [CrossRef]
  23. Takahashi, A.; Sakai, M.; Kato, T. Melting Temperature of Thermally Reversible Gel. VI. Effect of Branching on the Sol–Gel Transition of Polyethylene Gels. Polym. J. 1980, 12, 335–341. [Google Scholar] [CrossRef]
  24. Byrne, P.; Lloyd, G.O.; Applegarth, L.; Anderson, K.M.; Clarke, N.; Steed, J.W. Metal-induced gelation in dipyridylureas. New J. Chem. 2010, 34, 2261–2274. [Google Scholar] [CrossRef]
  25. Song, S.; Wang, L.; Yao, C.; Wang, Z.; Xie, G.; Tao, X. Crystallization of Sulfathiazole in Gel: Polymorph Selectivity and Cross-Nucleation. Cryst. Growth Des. 2020, 20, 9–16. [Google Scholar] [CrossRef]
  26. Ran, X.; Shi, L.; Zhang, K.; Lou, J.; Liu, B.; Guo, L. The Gelation Ability and Morphology Study of Organogel System Based on Calamitic Hydrazide Derivatives. J. Nanomater. 2015, 2015, 357875. [Google Scholar] [CrossRef] [Green Version]
  27. Malkin, A.Y.; Isayev, A.I. Rheology: Concepts, Methods and Applications; Chemtech Publishing: Toronto, ON, Canada, 2006. [Google Scholar]
  28. Jin, C.; Song, W.J.; Liu, T.; Xin, J.N.; Hiscox, W.C.; Zhang, J.W.; Liu, G.F.; Kong, Z.W. Temperature and pH Responsive Hydrogels Using Methacrylated Lignosulfonate Cross-Linker: Synthesis, Characterization, and Properties. ACS Sustain. Chem. Eng. 2018, 6, 1763–1771. [Google Scholar] [CrossRef]
  29. Khaltab, T.A.; Tiu, B.D.B.; Adas, S.; Bunge, S.D.; Advincula, R.C. pH triggered smart organogel from DCDHF-Hydrazone molecular switc. Dyes Pigment. 2016, 130, 327–336. [Google Scholar] [CrossRef] [Green Version]
  30. Matsuo, K.; Matsuoka, M. Solid-State Polymorphic Transition of Theophylline Anhydrate and Humidity Effect. Cryst. Growth Des. 2007, 7, 411–415. [Google Scholar] [CrossRef]
  31. Seton, L.; Khamar, D.; Bradshaw, I.J.; Hutcheon, G.A. Solvent-Mediated Central Metals Transformation from a Tetranuclear NiII Cage to a Decanuclear CuII “Pocket”. Cryst. Growth Des. 2010, 10, 3879–3886. [Google Scholar] [CrossRef]
  32. Patel, M.K.; Tailor, S.M.; Patel, V.H. DFT study and Hirshfeld surface analysis of third polymorph of sulfamerazine. AIP Conf. Proc. 2016, 1728, 020257. [Google Scholar] [CrossRef]
  33. Munroe, A.; Rasmuson, A.C.; Modnett, B.K.; Croker, D.M. Relative Stabilities of the Five Polymorphs of Sulfathiazole. Cryst. Growth Des. 2012, 12, 2825–2835. [Google Scholar] [CrossRef]
  34. Anwar, J.; Tarling, S.E.; Barnes, P. Polymorphism of Sulfathiazole. J. Pharm. Sci. 1989, 78, 337–342. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, K.; Fellah, N.; López-Mejías, V.; Ward, M.D. Polymorphic Phase Transformation Pathways under Nanoconfinement: Flufenamic Acid. Cryst. Growth Des. 2020, 20, 7098–7103. [Google Scholar] [CrossRef]
  36. Saha, S.; Desiraju, G.R. Acid···Amide Supramolecular Synthon in Cocrystals: From Spectroscopic Detection to Property Engineering. J. Am. Chem. Soc. 2018, 140, 6361–6373. [Google Scholar] [CrossRef] [PubMed]
  37. Surov, A.O.; Terekhova, I.V.; Bauer-Brandl, A.; Perlovich, G.L. Thermodynamic and Structural Aspects of Some Fenamate Molecular Crystals. Cryst. Growth Des. 2009, 9, 3265–3272. [Google Scholar] [CrossRef]
  38. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  39. Sheldrick, G.M. A short history of SHELX. Acta. Cryst. 2008, A64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  40. López-Mejías, V.; Kampf, J.W.; Matzger, A.J. Nonamorphism in flufenamic acid and a new record for a polymorphic compound with solved structures. J. Am. Chem. Soc. 2012, 134, 9872–9875. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of the possible G13 gelators.
Scheme 1. Synthesis of the possible G13 gelators.
Gels 09 00026 sch001
Figure 1. (a) Gels of G1 at 2% wt. in (from left to right) dichloromethane, ethanol, water, 1-pentanol, 1-propanol, 2-butanol, acetone, acetonitrile and ethyl acetate. (b) Gels of G1 at 1% wt. in (from left to right) dichloromethane, 1-pentanol, ethyl acetate, water and 1-propanol.
Figure 1. (a) Gels of G1 at 2% wt. in (from left to right) dichloromethane, ethanol, water, 1-pentanol, 1-propanol, 2-butanol, acetone, acetonitrile and ethyl acetate. (b) Gels of G1 at 1% wt. in (from left to right) dichloromethane, 1-pentanol, ethyl acetate, water and 1-propanol.
Gels 09 00026 g001
Figure 2. (a) Crystals of G1 from ethylene glycol. (b) Crystal packing in G1 showing hydrogen-bonded interactions between N-H and C=O groups (N···O 2.92 Å) [18].
Figure 2. (a) Crystals of G1 from ethylene glycol. (b) Crystal packing in G1 showing hydrogen-bonded interactions between N-H and C=O groups (N···O 2.92 Å) [18].
Gels 09 00026 g002
Figure 3. (a) Crystals of G2 from ethanol. (b) Crystal packing in the dihydrate of G2 crystallized from ethanol (N···O 2.9905(18), Owater···Owater 2.7672(13); Owater···O=C 2.7830(17) Å).
Figure 3. (a) Crystals of G2 from ethanol. (b) Crystal packing in the dihydrate of G2 crystallized from ethanol (N···O 2.9905(18), Owater···Owater 2.7672(13); Owater···O=C 2.7830(17) Å).
Gels 09 00026 g003
Figure 4. (a) SEM images for G1 xerogels dried from dichloromethane. (b) SEM images for compound G3.
Figure 4. (a) SEM images for G1 xerogels dried from dichloromethane. (b) SEM images for compound G3.
Gels 09 00026 g004
Figure 5. (a) Frequency sweep experiment at a concentration of 2% in acetonitrile. Stress sweep experiments at a concentration of 2% wt. for (b) G1 in 1-pentanol (σ = 1000 Pa) and (c) G1 in ethanol (σ = 1995 Pa).
Figure 5. (a) Frequency sweep experiment at a concentration of 2% in acetonitrile. Stress sweep experiments at a concentration of 2% wt. for (b) G1 in 1-pentanol (σ = 1000 Pa) and (c) G1 in ethanol (σ = 1995 Pa).
Gels 09 00026 g005
Figure 6. pH response of G1.
Figure 6. pH response of G1.
Gels 09 00026 g006
Figure 7. Chemical structures of the APIs chosen for crystallization in gels of G1.
Figure 7. Chemical structures of the APIs chosen for crystallization in gels of G1.
Gels 09 00026 g007
Figure 8. Crystals of flufenamic acid grown in dichloromethane gel of G1 (left) and with solution evaporation (right).
Figure 8. Crystals of flufenamic acid grown in dichloromethane gel of G1 (left) and with solution evaporation (right).
Gels 09 00026 g008
Table 1. CGC values for derivative G1.
Table 1. CGC values for derivative G1.
SolventCGC (%wt.)
Dichloromethane0.6
Ethanol1.5
1-Pentanol0.8
1-Propanol0.8
2-Butanol1.3
Acetone1.2
Acetonitrile1.5
Water1.0
Ethyl acetate0.6
Table 2. Tsol for gels of G1 at 2% wt. and 1% wt.
Table 2. Tsol for gels of G1 at 2% wt. and 1% wt.
SolventTsol (°C) (2% wt.)Tsol (°C) (1% wt.)
Dichloromethane6670
Ethanol82---
1-Pentanol11078
1-Propanol6263
2-Butanol105---
Acetone94---
Acetonitrile93---
Water12172
Ethyl acetate13274
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Torres-Moya, I.; Sánchez, A.; Saikia, B.; Yufit, D.S.; Prieto, P.; Carrillo, J.R.; Steed, J.W. Highly Thermally Resistant Bisamide Gelators as Pharmaceutical Crystallization Media. Gels 2023, 9, 26. https://doi.org/10.3390/gels9010026

AMA Style

Torres-Moya I, Sánchez A, Saikia B, Yufit DS, Prieto P, Carrillo JR, Steed JW. Highly Thermally Resistant Bisamide Gelators as Pharmaceutical Crystallization Media. Gels. 2023; 9(1):26. https://doi.org/10.3390/gels9010026

Chicago/Turabian Style

Torres-Moya, Iván, Abelardo Sánchez, Basanta Saikia, Dmitry S. Yufit, Pilar Prieto, José Ramón Carrillo, and Jonathan W. Steed. 2023. "Highly Thermally Resistant Bisamide Gelators as Pharmaceutical Crystallization Media" Gels 9, no. 1: 26. https://doi.org/10.3390/gels9010026

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