Next Article in Journal
A Visualized Microstructure Evolution Model Integrating an Analytical Cutting Model with a Cellular Automaton Method during NiTi Smart Alloy Machining
Previous Article in Journal
A Visual Representation for Accurate Local Basis Set Construction and Optimization: A Case Study of SrTiO3 with Hybrid DFT Functionals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization, and Analysis of Probenecid and Pyridine Compound Salts

1
National Engineering Research Center of Industrial Crystallization Technology, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
2
Collaborative Innovation Center of Chemical Science and Engineering, Tianjin 300072, China
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(7), 670; https://doi.org/10.3390/cryst14070670
Submission received: 25 June 2024 / Revised: 15 July 2024 / Accepted: 17 July 2024 / Published: 22 July 2024
(This article belongs to the Section Crystal Engineering)

Abstract

:
This study aimed to address the issue of the low solubility in the model drug probenecid (PRO) and its impact on bioavailability. Two salts of probenecid (PRO), 4-aminopyridine (4AMP), and 4-dimethylaminopyridine (4DAP) were synthesized and characterized by PXRD, DSC, TGA, FTIR, and SEM. The crystal structures of the two salts were determined by SCXRD, demonstrating that the two salts exhibited different hydrogen bond networks, stacking modes, and molecular conformations of PRO. The solubility of PRO and its salts in a phosphate-buffered solution (pH = 6.8) at 37 °C was determined, the results showed that the solubility of PRO salts increased to 142.83 and 7.75 times of the raw drug, respectively. Accelerated stability experiments (40 °C, 75% RH) showed that the salts had good phase stability over 8 weeks. Subsequently, Hirshfeld surface (HS), atom in molecules (AIM), and independent gradient model (IGM) were employed for the assessment of intermolecular interactions. The analyses of salt-forming sites and principles were conducted using molecular electrostatic potential surfaces (MEPs) and pKa rules. The lattice energy (EL) and hydration-free energy (EHF) of PRO and its salts were calculated, and the relationships between these parameters and melting points and the solubility changes were analyzed.

1. Introduction

The synthesis of the multi-component crystals (such as cocrystals, salts, and solvates) [1] of active pharmaceutical ingredients (APIs) in order to enhance their physicochemical properties, including solubility, dissolution rate, bioavailability, stability, and hygroscopicity, has received increasing attention and found widespread application in the pharmaceutical industry [2,3]. Notably, 40% of currently marketed drugs exhibit poor water solubility [4], and the formation of salts is the preferred approach for addressing the low solubility of pharmaceutical compounds [5]. According to the IUPAC definition, a salt is “the substance formed by the combination of cations and anions”. The formation of a drug molecule as a salt requires the presence of an ionized API along with another counterion. Therefore, for drugs with ionization ability, synthesizing salts represents a highly established method for enhancing their physicochemical properties [6]. The formation of salt derivatives has been shown to significantly enhance the solubility, dissolution rate [7], stability [8,9,10], and bioavailability [11] of pharmaceutical compounds while also extending the patent exclusivity period for marketed drugs. It is estimated that more than 50% of drug molecules are administered in the form of salts [12].
Probenecid (PRO; Figure 1) (Chemical name: p-[(dipropyl amino) sulfonyl] benzoic acid, molecular formula: C13H19NO4S, CAS: 57-66-9), a synthetic sulfonamide with the dual effects of promoting uric acid excretion and inhibiting penicillin excretion, has long been used in the treatment of chronic gout [13,14]. The latest research has demonstrated that bumetanide effectively modulates human renal physiological functions by inhibiting ATP transporter proteins in the renal collecting tubules and proximal tubules. Additionally, its potential to mitigate transient global cerebral ischemia/reperfusion injury in mice underscores its significant pharmaceutical value and wide-ranging applications [15,16]. However, according to the biopharmaceutical classification system (BCS) [17,18], PRO is a BCS II drug with good permeability but poor water solubility. According to the literature reports, its solubility in water at 37 °C is only 72.2 μg/mL [19], which influences its bioavailability. Although the water solubility of the PRO is significantly limited, there has been limited progress in terms of developing methods for synthesizing salts to enhance the water solubility of PRO. Recently, there have been studies on the elastic crystals formed by the coprecipitation of PRO with 4,4-azopyridine [20,21], 4,4-bipyridine [22], azacytidine, and piperazine [23]. Additionally, research has shown an enhancement in solubilitiy through the coprecipitation of PRO with benzamide [19,24] and 1,2-bis(4-pyridyl) ethene (2.43 times increase) [25]. Furthermore, there have been investigations into the N, N-dimethylformamide solvate and the pyridine solvate of PRO [22,23]. Due to the presence of a carboxyl group that can dissociate in PRO molecules, we aim to enhance its solubility through the synthesis of PRO salts. This study computed and analyzed the 3D full interaction map [26] (FIM) of PRO and selected pyridine-based basic compounds that are prone to proton transfer and form salts for experimental screening. Subsequently, we carried out experimental screening of the coformers by means of a liquid-assisted grinding approach, which is more efficient than conventional solvent-free mechanical grinding [27,28], and characterized them using powder X-ray diffraction (PXRD) and differential scanning calorimetry (DSC). We finally found that PRO can form a new phase with 4-aminopyridine (4AMP) and 4-dimethylaminopyridine (4DAP). As commonly basic coformers [29,30], 4AMP and 4DAP have been employed in the formation of salts with Furantoin [31], Piroxicam and Meloxicam [32], Ibuprofen [33], and other drugs to enhance the solubility and other physicochemical properties of these pharmaceutical compounds.
In this investigation, PRO was chosen as the model drug, and its salts were synthesized with coformers 4-aminopyridine (4AMP, CAS: 504-24-5) (PRO)(4AMP)+ (1:1) and 4-dimethylaminopyridine (4DAP, CAS:1122-58-3) (PRO)(4DAP)+ (1:1) (Figure 1). Moreover, the salts were comprehensively characterized by powder X-ray diffraction (PXRD), differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), Fourier transform infrared spectroscopy (FTIR), and scanning electron microscopy (SEM). The results revealed distinct PXRD patterns, FTIR spectra, melting points, decomposition temperatures, and crystal morphologies for the salts compared to the raw materials. The crystal structures of the salts were determined by single crystal X-ray diffraction (SCXRD), revealing distinct hydrogen bond networks, packing modes, and PRO molecular conformations. Given the primary absorption site of PRO in the small intestine, equilibrium solubility data for PRO and its salts were obtained in a phosphate-buffered solution (pH = 6.8) at 37 °C to simulate intestinal conditions. Results indicated that the solubility of (PRO)(4AMP)+ and (PRO)(4DAP)+ increased to 142.83 and 7.75 times, respectively, compared to the raw drug, significantly improving the poor solubility of PRO. Accelerated stability experiments were conducted on the salts at 40 °C and 75% relative humidity (RH), and no observable crystalline changes were observed in either salt within 8 weeks, indicating excellent phase stability. The molecular interactions within the salts were analyzed using Hirshfeld surface (HS), atom in molecules (AIM), and independent gradient model (IGM). The salt-forming sites and principles were analyzed using molecular electrostatic potential surfaces (MEPs) and pKa rules. The lattice energy (EL) and hydration-free energy (EHF) of PRO and its salts were calculated, and the relationships between these parameters and melting points and solubility changes were analyzed. It is particularly necessary to explain that 4AMP and 4DAP have a certain toxicity and are not suitable for human medication. However, the focus of this study is concept verification in regard to improving the solubility of soluble drugs through salt formation, which can provide new ideas for improving the solubility of drugs by designing salt formation.

2. Materials and Methods

2.1. Materials

The PRO (purity: 98.0%) and coformers (purity: ≥98.0%) used in this study were purchased from Shanghai Dibai Biotechnology Co., Ltd., Shanghai, China and Tianjin Xiensi Biotechnology Co., Ltd., Tianjin, China. All solvents were of analytical grade and provided by Tianjin Jiangtian Chemical Technology Co., Ltd., Tianjin, China. All chemicals can be directly used without further purification. Deionized water is prepared in the laboratory, as shown in Table S1.

2.2. Screening of Salts

The 3D full interaction map (FIM) of PRO were computed and generated by importing the .CIF format files of the PRO’s parent crystal structure (CCDC No: 987956) into Mercury 4.2.0 software (Cambridge Crystallographic Data Centre, Cambridge, UK) (Figure 2a) [34], which shows the potential hydrogen bond donor and acceptor sites in the PRO molecule. The dark outlined (red/blue) regions indicate a higher tendency for synthon formation, while the transparent regions have a lower tendency. The PRO molecule contains hydrogen bond donors and acceptors, such as carboxylic acid and sulfonamide groups (Figure 2b), that can form salts with basic compounds containing pyridine rings or other nitrogen-containing heterocycles [35] (Figure 2c). The coformers selected in this experiment include 4-aminopyridine (4AMP), 4-dimethoxypyridine (4DAP), 4-hydroxypyridine (4HDP), 2-pyridinecarboxamide (2PCA), 3-aminopyridine (3AMP), and 2,6-diaminopyridine (26DAP).
Experimental screening of the coformers was conducted using liquid-assisted grinding (LAG). An amount of 1 mmol of PRO and 1 mmol of the coformer were weighed using an analytical balance, then added to an agate mortar. Subsequently, 40 μL of ethanol was added in four equal portions, and the sample was dried at 50 °C for 24 h after grinding for 1 h. Finally, PXRD and DSC analyses were employed for characterization, as depicted in Figures S1 and S2. From these analyses, it can be observed that new peaks emerged in the PXRD patterns of PRO after grinding with 4AMP and 4DAP, and the DSC curves indicated the existence of new melting points. Consequently, we are convinced that PRO formed two new phases with 4AMP and 4DAP.

2.3. Synthesis of Salts

2.3.1. Synthesis of (PRO)(4AMP)+

The PRO (285.4 mg, 1 mmol) and 4AMP (94.2 mg, 1 mmol) were weighed and added to 10 mL of acetone. The slurry was suspended at room temperature for 12 h, filtered, and then dried at 50 °C for 24 h to yield a white solid powder.
Due to the formation of oil during the slow solvent evaporation method for single crystal preparation, single crystal was obtained using the slow cooling crystallization method. PRO (57.1 mg, 0.2 mmol) and 4AMP (18.8 mg, 0.2 mmol) were weighed and added to 10 mL of acetone, the mixture was heated to dissolve at 50 °C before being cooled at a rate of 5 °C/h until colorless rod-shaped single crystals formed at a temperature of 20 °C.

2.3.2. Synthesis of (PRO)(4DAP)+

The PRO (285.4 mg, 1 mmol) and 4DAP (122.2 mg, 1 mmol) were weighed and added to 10 mL of ethyl acetate. The slurry was then suspended at room temperature for 12 h, filtered, and then dried at 50 °C for 24 h to yield a white solid powder.
Single crystal was obtained using the slow solvent evaporation method. Additionally, PRO (28.6 mg, 0.1 mmol) and 4DAP (12.3 mg, 0.1 mmol) were dissolved in ultrasonicated ethyl acetate and slowly evaporated to obtain colorless rod-shaped single crystals after several days.

2.4. Characterization

2.4.1. Single Crystal X-ray Diffraction (SCXRD)

SCXRD measurement was conducted on a diffractometer (Saturn 70CCD, Rigaku, Japan) using Mo Kα radiation (λ = 0.71073 Å). The integration and scaling of the intensity data were performed by the CrysAlisPRO 1.171.39.46 program (Rigaku Oxford Diffraction, 2018). The structure was solved using direct methods in SHELXT [36] and refined using the full matrix least-squares method in SHELXL [37], both of which were performed under OLEX2-1.2 [38]. The non-H atoms were refined using anisotropic parameters. The riding model with Uiso = 1.2 − 1.5Ueq(C) was employed for the geometric localization and refinement of H atoms on C atoms, while the H atoms bonded to N and O atoms were located based on the residual peaks of electron density in the Fourier maps and refined with isotropic parameters. Absorption effect data were corrected using SADABS [39] and hydrogen bond data were detected using PLATON [40]. Mercury 4.2.0 software was used for visualization of the crystal structure [41].

2.4.2. Powder X-ray Diffraction (PXRD)

PXRD was measured using a powder X-ray diffractometer (MiniFlex600, Rigaku, Japan) with Cu radiation (λ = 1.54178 Å). The working voltage and current were 40 kV and 100 mA, respectively. The samples were measured within the 2θ range of 2–35° with a step size of 0.01° and a scanning rate of 10°/min. The data were collected at room temperature. The simulated PXRD patterns were obtained using Mercury 4.2.0 software.

2.4.3. Differential Scanning Calorimetry (DSC)

DSC was performed on a Mettler DSC 1 STARe system (Mettler–Toledo, Greifensee, Switzerland) that was calibrated with indium standards prior to analysis. A sample of 5–10 mg was heated in a standard aluminum crucible with a needle hole in a nitrogen atmosphere at a heating rate of 10 °C/min and a nitrogen flow of 70 mL/min. An empty aluminum crucible with a needle hole was used as the reference.

2.4.4. Thermogravimetric Analysis (TGA)

TGA was performed using the Mettler TGA/DSC 1 STARe system (Mettler–Toledo, Greifensee, Switzerland) under a nitrogen atmosphere with a flow rate of 40 mL/min. Each analysis utilized approximately 5–10 mg of a sample placed in an alumina crucible. The standard uncertainty was ±0.00001 g and the sample heating rate was set at 10 °C/min.

2.4.5. Fourier Transform Infrared Spectroscopy (FT-IR)

The ALPHA II infrared instrument (Bruker, Ettlingen, Germany) equipped with an ATR attachment was used to collect infrared spectra in the scanning range of 4000 to 400 cm−1 under ambient conditions. At least 16 spectra were collected for each sample, and the average value was taken. The instrument resolution was 4 cm−1.

2.4.6. Scanning Electron Microscopy (SEM)

To analyze the crystal morphology of the samples, images of the samples were recorded using a scanning electron microscope (SEM, TM3000, Hitachi Corporation, Tokyo, Japan).

2.4.7. Equilibrium Solubility Measurement

The equilibrium solubility of PRO and its salts in a phosphate-buffered solution (pH = 6.8, simulating intestinal fluid) was determined at 37.0 ± 0.1 °C using the shake flask method. The phosphate-buffered solution (pH = 6.8) was prepared according to the Chinese Pharmacopoeia (2015 edition) [42]. An excess sample was suspended in the buffer solution at 37.0 ± 0.1 °C for 48 h to achieve equilibrium, followed by filtration of the supernatant through a PTFE filter membrane (0.45 μm) and appropriate dilution. The liquid chromatography peak areas of a series of PRO solutions with known concentrations ranging from 10 to 400 μg/mL were determined. Subsequently, with the concentrations and chromatographic peak areas serving as the abscissa and ordinate, respectively, the least square method was employed to linearly fit the scattered data, thereby obtaining the working curve for solubility testing. The experiment were performed using high-performance liquid chromatography (Agilent 1220, Shanghai, China) under specific test conditions, including an Agilent Extend C18 Column (250 × 4.6 mm, 5 μm), a column temperature of 30 °C, a UV detection wavelength of 245 nm, a methanol–water mobile phase (75:25), and a flow rate of 1.0 mL/min [43]. The PXRD analysis was employed for monitoring phase transitions. The equilibrium solubility experiment was repeated three times for accuracy.

2.4.8. Accelerated Stability Experiment

In order to test the stability of PRO and its salts, accelerated stability experiments were conducted in a medicine stability test chamber (HWS-70B, Taisite Instrument Co., Ltd., Tianjin, China) at 40 ± 1 °C and 75 ± 1.5% RH. After 8 weeks, the samples were removed from the chamber and analyzed using PXRD to evaluate their phase stability.

2.5. Computational Details

2.5.1. Hirshfeld Surface (HS)

The Hirshfeld surface (HS) and two-dimensional (2D) fingerprint diagram were computed and generated by importing the .CIF format files of the PRO and its salts into CrystalExplorer 17.5 software for a clear and intuitive study of molecular interactions within the crystal structure as well as for quantitative analysis [44,45].

2.5.2. Atom in Molecules (AIM) and Independent Gradient Model (IGM)

The crystals structures of PRO [34] and its salts were optimized using the CASTEP module of Materials Studio 6.0 (Accelrys, San Diego, CA, USA) [46]. Structural relaxation of the positions of all atoms was performed by plane wave density functional calculations, the cell parameters were fixed during the optimization process. The Perdew–Burke–Ernzerhof (PBE) generalized gradient approximation (GGA) was used as an exchange–correlation density functional [47] along with ultra-soft pseudopotential [48] plus Grimme’s D2 [49] dispersion correction. A cut-off energy of 780 eV was set, and a Brillouin zone integration was performed using discrete 2 × 2 × 2 k-point sampling of the original cell. For structural relaxation, thresholds for energy, force, and atomic shift were set to 10−5 eV, 5 × 10−2 eV/A, and 10−3 A, respectively. After geometry optimization, molecular clusters required for calculations were extracted from crystal supercells. Single point energy levels of the molecular cluster were calculated at the RI [50]-wB97M-V [51]/def2-TZVP [52] level using ORCA 4.2.0 [53] to generate wave functions for AIM [54] and IGM [55] analysis.

2.5.3. Molecular Electrostatic Potential Surfaces (MEPs)

The geometric optimization of raw compounds was conducted using ORCA 4.2.0 software at the RI-B3LYP-V/def2-TZVP level while obtaining wavefunction files for both the molecule and salts to facilitate analysis of molecular electrostatic potential surfaces (MEPs). Subsequently, visualization of the MEPs, AIM, and IGM was achieved using Multiwfn 3.8 [56,57,58] and VMD 1.9.3 [59].

2.5.4. pKa

The pKa values of PRO and coformers were computed utilizing MarvinSketch (ChemAxon, Budapest, Hungary) [60,61].

2.5.5. Lattice Energy (EL) and Hydration-Free Energy (EHF)

In this study, the Forcite module under Materials Studio 6.0 (Accelrys, San Diego, CA, USA) was utilized to compute the lattice energy (EL) for assessing the lattice strength of PRO and its salts. The geometric optimization of the lattice was accomplished with the Compass force field method [62,63], and the calculation of EL can be shown in Equation (1) as follows:
EL = E(bulk)/Z − EAEB
where Z represents the number of asymmetric units, E(bulk) represents the energy of a crystal unit cell, and EA and EB represent the relaxed energy of A and B molecules (kcal/mol), respectively.
The uESE method is a simple, efficient and accurate method for calculating the solvation free energy of ions and neutral molecules. For anionic systems in particular, the uESE method is more convenient and accurate than the traditional SMD method [64]. The specific procedural steps involve geometric optimization of cations, anions, and PRO molecules at the RI-B3LYP-D3(BJ)/def2-TZVP level using ORCA 4.2.0, followed by inputting the optimized structures into Multiwfn 3.8 software to generate calculation files for the uESE input file. Finally, the uESE program is utilized for computing hydration-free energy.

3. Results and Discussion

3.1. Crystal Structure Analysis

3.1.1. Single Crystal X-ray Diffraction Analysis

The single crystal structures of (PRO)(4AMP)+ and (PRO)(4DAP)+ were determined through SCXRD analysis, with the corresponding crystallographic data and refined details being presented in Table 1. These specific datasets have been deposited in the Cambridge Crystallographic Data Centre under CCDC numbers 2283803 and 2360637. Additionally, the C-O bond length data for the carboxyl group of PRO salts were summarized in Table 2. Table 3, Tables S2 and S3 contain information on hydrogen bonds and other weak interactions information within the salts’ crystal structure.
The bond length of carboxyl groups in single crystal data is frequently utilized for distinguishing between cocrystals and salts. The disparity in bond length between the two C-O bonds within the carboxyl group, denoted as ΔDC-O, serves as a crucial metric for discerning proton transfer. A smaller ΔDC-O value signifies the formation of a salt via proton transfer, whereas a larger value (>0.08 Å) indicates cocrystal formation without proton transfer [65]. Both single crystal structures in this study exhibit relatively small ΔDC-O values (Table 2), corroborating proton transfer and confirming the formation of salt rather than cocrystals.
(PRO)(4AMP)+ salt (1:1) The crystal structure of (PRO)(4AMP)+ salt crystallizes in the orthorhombic crystal system with a space group of P212121 (Z = 4). The asymmetric unit consists of one (PRO) anion and one (4AMP)+ cation. In the asymmetric unit, the H atom on the carboxyl group of the PRO molecule transfers to the N atom on the pyridine ring of the 4AMP molecule, forming the salt. Since the benzene ring of the (PRO) anion and the pyridine ring of the (4AMP)+ cation were reversed by 36.43° (Figure S3), no cyclic hydrogen bond motif was formed. The (PRO) anion and (4AMP)+ cation are connected through a charge-assisted hydrogen bond N2+-H2⋯O2 (1.885 Å; 167.36°) (Figure 3a). The asymmetric units are connected into a 1D chain through hydrogen bond C12-H12B⋯O3 (2.585 Å; 127.55°; 1 + x, y, z) interactions (Figure 3b). The 1D chains are further connected into a double 1D chain through hydrogen bond (C14-H14⋯O4 (2.529 Å; 129.62°; -1/2 + x, 1/2 − y, 1 − z) and C18-H18⋯O4 (2.556 Å; 144.50°; 1/2 + x, 1/2 − y, 1 − z)) (Figure 3c). In addition, weak interactions, such as C7-H7⋯π (2.721 Å, 148.95°; 1/2 + x, 1/2 − y, 1 − z), and C13-H13A⋯π (2.862 Å, 154.75°; 1/2 + x, 1/2 − y, 1 − z) (Figure 3c), play a crucial role in the crystal structure. Observed from the a-axis direction, the double chains are connected into a 3D structure through hydrogen bond (N3-H3B⋯O2 (2.042 Å; 167.51°; 1 − x, −1/2 + y, 3/2 − z) and N3-H3A⋯O1 (1.965 Å; 167.77°; −x, −1/2 + y, 3/2 − z)) (Figure 3d).
(PRO)(4DAP)+ salt (1:1) The crystal structure of (PRO)(4DAP)+ salt crystallizes in the monoclinic P21/c (Z = 4) space group, with one (PRO) anion and one (4DAP)+ cation in the asymmetric unit (Figure 4a). There is overall disorder at the terminal propyl group of PRO, with PART1 accounting for 0.585 (6), and the crystal structure is resolved using the larger PART1 fraction. Proton transfer is detected from the carboxyl group of PRO to the pyridine N atom of 4DAP, leading to the formation of salt. The phenyl ring of the (PRO) anion and the pyridine ring of the (4DAP)+ cation exhibit a slight tilt, with only a 5.26° rotation occurring between their planes (Figure S3), so (PRO) anions and (4DAP)+ cations form a cyclic motif R 2 2 (7) through hydrogen bond C18-H18⋯O2 (2.356 Å; 129.74°) and charge assisted hydrogen bond N2+-H2⋯O1 (1.858 Å; 172.85°). Observed from the b-axis, the asymmetric units are connected by hydrogen bonds C14-H14⋯O2 (2.350 Å; 154.29°; −1 + x, y, z) to form 1D chains (Figure 4b). Then, observed from the a-axis, the 1D chains are connected by C10-H10C⋯O4 (2.573 Å; 154.49°; x, 1 + y, z) to form 2D planes (Figure 4c). The 2D planes are connected by C20-H20A⋯π (2.653 Å; 148.86°; 1 − x, 1 − y, 1 − z) and π⋯π (3.800 Å; 0.03(9)°; −x, 2 − y,1 − z) weak interactions to form 2D double-layer planes. The centroid–centroid distance within the π⋯π stacking amounts to 3.800 Å; the plane–plane distance is 3.448 Å; the angle between the normal of the plane and the centroid connection is 24.90°; the slippage is 1.597 Å. According to Christoph Janiak et al., the aforementioned π⋯π stacking represents a typical parallel-displaced π⋯π stacking of rings and encompasses the contribution of π⋯σ attraction [66]. Subsequently, the 2D double-layer planes are interconnected along the c-axis via weak interlayer interactions, thereby forming a complete 3D crystal structure (Figure 4d).

3.1.2. Conformation and Packing Similarity Analysis

The flexibility of PRO molecular conformation can significantly impact the assembly of multi-component crystals. In this investigation, we examined the diverse conformations of PRO in both its parent crystal [34] and salt crystals, followed by overlaying and analyzing the conformations using Mercury 4.2.0 software (Figure 5a). It is evident from the data that the terminal dipropyl amino group is particularly prone to distortion within the PRO molecule. Furthermore, proton transfer during salt formation induces specific alterations in PRO conformation. Notably, compared to the PRO parent crystal, the conformation of PRO in the (PRO)(4AMP)+ salt crystal exhibits a certain degree of resemblance; conversely, significant differences are observed between the conformation of PRO in the (PRO)(4DAP)+ salt crystal and that in the parent crystal. Specifically, within the (PRO)(4DAP)+ salt structure, rotation of the carboxyl group occurs relative to both (PRO)(4AMP)+ salt and pure PRO crystalline forms; consequently, these two salts are formed through charge-assisted hydrogen bonds N2+-H2⋯O1 and N2+-H2⋯O2 (Figure S3).
The molecular packing similarity of two salts was further evaluated using Mercury 4.2.0 software, taking into consideration various structural and ensemble factors. Comparison of the analysis of seven molecules revealed that only one molecule exhibited similar packing positions (Figure 5b). This suggests a poor similarity in crystal packing among them, possibly due to variances in the hydrogen bond networks of the two salts (Figure 5c–d).

3.2. Powder Characterization Analysis

3.2.1. Powder X-ray Diffraction Analysis

The experiment involved the preparation of powder samples of (PRO)(4AMP)+ and (PRO)(4DAP)+ salts using the slurry suspension method, followed by their characterization using PXRD. The single crystal structures obtained from SCXRD for the salt crystals were utilized to simulate the PXRD pattern of the salts, as shown in Figure 6. It is evident from the figure that the characteristic peaks of the obtained salts are distinctly discernible from those of the initial raw materials, thereby confirming the formation of new high-purity phases rather than physical mixtures. Furthermore, there is a fundamental consistency between the experimental PXRD patterns and simulated PXRD patterns based on a single crystal structure, thus substantiating both representativeness and high crystallinity.

3.2.2. Thermal Analysis

DSC is a commonly used characterization method for analyzing the thermal behavior of drugs. The formation of salts can change the original chemical structure and cause changes in thermal properties, so the generation of new phases can be determined by the differences in thermal characteristics during drug heating. The DSC curves of the raw materials and salts are shown in Figure 7, and the results show that the melting point (Tonset) of PRO, 4AMP, and 4DAP are 198.73 °C, 157.42 °C, and 111.75 °C, respectively. The experimental melting point of (PRO)(4AMP)+ is 132.04 °C, which is lower than the melting point of raw materials. However, the melting point of (PRO)(4DAP)+ is 141.54 °C, which is between the melting points of the two starting materials. The melting enthalpies of (PRO)(4AMP)+ and (PRO)(4DAP)+ are −9.57 kcal/mol and −10.51 kcal/mol, respectively. By analyzing the results together with the PXRD results, it was shown that the formation of salts, rather than physical mixtures, was responsible for the new phase.
TGA is a widely utilized method for thermal analysis, providing valuable insights into the thermal stability and solvent composition of pharmaceuticals. The TGA-DSC curves in Figure 8 reveal that (PRO)(4AMP)+ salt exhibits no weight loss prior to decomposition, possesses a melting point lower than the decomposition temperature, and demonstrates robust thermal stability. Conversely, (PRO)(4DAP)+ salt undergoes simultaneous melting and decomposition, indicating inferior thermal stability. These observations may be attributed to the lower boiling point of 4DAP (162 °C) compared to that of 4AMP (273 °C). Additionally, the absence of residual solvents (including water) in the medicine salts is evident from the figure.

3.2.3. Spectral Analysis

In order to gain deeper insights into the salt formation process, FTIR spectroscopy was employed to analyze the salts of PRO and their respective raw materials. The comparative results are presented in Figure S4, with detailed characteristic peak information being provided in Table S4. For pure PRO, the characteristic peaks were observed at 1686, 2964, 1284, and 1341 cm−1, corresponding to C=O stretching vibration, O-H stretching vibration, C-N stretching vibration, and S=O stretching vibration, respectively. As for coformers, the C=N stretching vibrations of 4DAP and 4AMP were identified at 2900 cm−1 and 3003 cm−1, respectively. No N-H vibration peak was detected in pure 4DAP; however, a peak appeared at 3072 cm−1 in the formed salt, indicating N-H stretching vibration and confirming proton transfer occurrence. A strong characteristic stretching vibration corresponding to NH2 was detected at 3432 cm−1 in pure 4AMP. Nevertheless, upon salt formation, the NH2 stretching vibration vanished, and, subsequently, a broad peak emerged within the range of 3362–3318 cm−1. These manifestations suggest that the amino groups have associated. In combination with the SCXRD analysis in Section 3.1.1, since the acidic proton forms a hydrogen bridge between the N atom on the 4AMP pyridine ring and the O atom of the PRO carboxyl group and the NH2 of 4AMP and the carboxyl group of PRO form strong N-H⋯O hydrogen bonds, these factors lead to the reduction of the stretching vibration intensity and frequency of NH2. Furthermore, the shift of the C=O peak to a lower frequency subsequent to salt formation further implies that the original hydrogen bond network is disrupted and substituted by a newly formed hydrogen bond network, thereby further validating the formation of the salt.

3.2.4. SEM Analysis

The prepared powders were subjected to SEM analysis to assess alterations in crystal morphology compared to the starting material (Figure 9). As depicted in the picture, PRO demonstrates elongated plate-like crystals, 4AMP displays block-shaped crystals, and 4DAP exhibits thick plate-like crystals (Figure 9a–c). Interestingly, the (PRO)(4AMP)+ salt manifests a sleek rod-like crystal structure on its surface (Figure 9d), whereas the (PRO)(4DAP)+ salt presents a rugged and fragmented plate-like crystal structure on its surface (Figure 9e). These phenomena imply that the salts synthesized experimentally exhibit distinct crystal morphologies compared to the raw materials.

3.2.5. Equilibrium Solubility Analysis

By determining the liquid chromatographic peak areas of a series of PRO solutions with known concentrations (10–400 μg/mL) and then taking the solution concentration C (μg/mL) as the abscissa and the chromatographic peak area A (mAU*s) as the ordinate, as well as performing linear fitting using the least squares method on the data, the working curve A = 40.4293 × C + 23.0174 for the solubility test was obtained, and the value of R2 is 0.9999. The working curve is shown in Figure S5. Moreover, since it is known that the main absorption site of PRO is in the small intestine, the equilibrium solubility data of PRO, (PRO)(4AMP)+ and (PRO)(4DAP)+ were measured in phosphate-buffered solution (pH = 6.8) at 37 °C, and the residual solids after dissolution experiment were filtered and analyzed using PXRD to determine the type of solid phase at equilibrium. The results showed that both salts retained their original crystal patterns after the dissolution experiment (Figure S6). PRO, (PRO)(4AMP)+ and (PRO)(4DAP)+ were 4.61 mg/mL, 658.45 mg/mL, and 35.71 mg/mL, respectively, in phosphate-buffered solution at 37 °C (pH = 6.8). Among them, the solubility of (PRO)(4AMP)+ and (PRO)(4DAP)+ increased to 142.83 and 7.75 times that of PRO, respectively, as shown in Table 4. The increase in solubility may be due to ionization and the reduction in lattice energy after salt formation, while water is a polar solvent and easily dissolves various ionic compounds.

3.2.6. Accelerated Stability Analysis

The stability of PRO and the salts (PRO)(4AMP)+ and (PRO)(4DAP)+ was assessed under accelerated conditions of 40 ± 1 °C and 75 ± 1.5% RH for 8 weeks (Figure S7). The PXRD patterns remained unchanged before and after the accelerated stability experiment, indicating high stability in the drug salts under elevated humidity without undergoing phase transformation or decomposition, which is crucial for sample storage.

3.3. Computational Analysis

3.3.1. HS Analysis

In order to visualize the molecular interactions within the multi-component crystals, the 3D Hirshfeld surfaces of (PRO)(4AMP)+ and (PRO)(4DAP)+ were computed using CrystalExplorer 17.5 software. The resulting 3D Hirshfeld surface is depicted in Figure 10a,b, with distinct colors denoting various interaction types: red signifies strong interactions, white denotes moderate interactions, and blue indicates negligible interactions. Analysis of the 3D Hirshfeld surface diagram reveals that the red regions are predominantly localized around the carboxyl group of the PRO molecule, signifying its a primary hydrogen bonding site involved in intermolecular interactions. In order to further investigate the weak interactions within the crystal, additional analysis of the molecular packing was performed, resulting in the generation of color mapping maps for shape index and curvedness on the Hirshfeld surface (Figure 10a,b). The red regions and blue regions in the shape index map correspond to depressions and protrusions, respectively. Alternating red and blue regions on the PRO surface indicate various weak interactions connecting PRO molecules with ligand molecules. Furthermore, analysis of the curvedness map reveals that the PRO molecule exhibits a broad, flat surface, suggesting planar stacking interactions on its sides (e.g., π⋯π interaction).
To further quantify the degree of molecular interactions, 2D fingerprint plots were computed for (PRO)(4AMP)+ and (PRO)(4DAP)+, with di and de denoting the distances from points on the Hirshfeld surface to the nearest nucleus inside and outside the surface, as depicted in Figure 10c,d. The predominant portion of the fingerprint spectrum corresponds to H-H contacts, while the protruding segment represents O-H contacts that contribute to O⋯H hydrogen bonds within the crystal structure. Due to nitrogen atoms in the pyrimidine ring being involved in salt formation, N-H contacts do not exhibit a protruding segment; instead, C-H contacts primarily contribute to another region.
Furthermore, the 2D fingerprint plots in Figure 10c,d illustrate varying proportions of interactions, with detailed data being provided in Table S5. In the case of salts (PRO)(4AMP)+ and (PRO)(4DAP)+, the predominant H-H contacts denote van der Waals interactions, constituting the highest percentage among all interactions. The 2D plane of (PRO)(4DAP)+ primarily forms a 3D structure through van der Waals forces, resulting in significantly larger H-H contacts (55.3%) compared to those of (PRO)(4AMP)+ (49.6%). Additionally, O-H contacts exhibit higher percentages (24.0% and 32.5%), signifying the crucial role of O⋯H hydrogen bonds in crystal packing. The C-H contacts (22.8%) observed in (PRO)(4AMP)+ are largely attributed to C-H⋯π weak interactions that contribute to structural stability. Both compounds have minimal N-H contacts (1.2% and 0.5%, respectively) as a result of nitrogen atoms participating in salt formation within the pyrimidine ring structure; however, there is a higher proportion of C-C contacts for (PRO)(4DAP)+ at 0.5%, indicating that the π⋯π weak interactions contributes to structural stabilization.

3.3.2. MEPs Analysis

Molecular electrostatic potential surfaces (MEPs) has been widely utilized for predicting molecular interaction sites and characterizing molecular recognition patterns within crystal structures. The MEP value reflects the strength of these interactions, which is crucial for understanding the formation of multi-component crystals. Figure 11a–c illustrates the MEPs of PRO, 4AMP, and 4DAP before salt formation, with orange denoting positive electrostatic potential and blue representing negative electrostatic potential. The peaks and troughs in the electrostatic potential are highlighted by yellow and blue spheres, respectively.
The global maximum electrostatic potential (+63.06 kcal/mol) for the PRO molecule is located at the O-H bond of the carboxyl group, indicating its propensity to donate a proton or act as a hydrogen bond donor. The primary hydrogen bond acceptor is the O atom of the carboxyl group, with an electrostatic potential of −33.10 kcal/mol. Conversely, the O atom of the sulfonyl group exhibits the global minimum electrostatic potential (−41.99 kcal/mol). According to the principle of electrostatic potential complementarity, the area with the highest electrostatic potential in PRO tends to be bind with the regions with the lowest electrostatic potential in 4AMP and 4DAP. Both 4AMP (−45.14 kcal/mol) and 4DAP (−47.06 kcal/mol) exhibit the lowest electrostatic potential at the N atom of the pyridine ring, thereby attracting the carboxyl group in PRO to facilitating proton transfer, leading to the formation of (N2+-H2⋯O2) and (N2+-H2⋯O1) salts. The H atom in the amino group of 4AMP exhibits the highest electrostatic potential (+43.35 kcal/mol), whereas in 4DAP, the H atom of the pyridine ring demonstrates a relatively higher electrostatic potential (+19.15 kcal/mol) and tends to form hydrogen bonds with the carboxyl group and sulfonic group of PRO.
Additionally, following the formation of salts, as shown in Figure 11d,e, the positive potential becomes concentrated on the 4AMP and 4DAP cations while the negative potential is primarily focused on the PRO anion due to proton transfer neutralizing the electrostatic potential in PRO, 4AMP, and 4DAP. After salt formation, PRO engages with 4AMP and 4DAP through a highly compatible electrostatic complementary interaction, thereby promoting overall system stability.

3.3.3. pKa Analysis

According to the ΔpKa empirical theory (pKa (base) − pKa (acid)), a cocrystal is formed when ΔpKa < 0 and a salt is formed when ΔpKa > 3. When 0 < ΔpKa < 3, either a salt or a cocrystal may be formed, which is in the intermediate state of proton transfer. The ΔpKa values of PRO, 4AMP, and 4DAP are 3.3 and 3.9, respectively, indicating that salts are formed, which is also confirmed by the SCXRD results (Table S6). However, in a recent study, Cruz-Cabeza re-examined the ΔpKa rule by statistically analyzing the dataset of 6465 crystal complexes and discovered that the region of the intermediate state of proton transfer is broader, ranging from −1 < ΔpKa < 4 [67]. Based on this, we can infer that the two new solid forms prepared in this experiment are in the intermediate state of proton transfer, being salt–cocrystal continua [68] rather than complete salts.

3.3.4. AIM Analysis

The purpose of AIM analysis is to investigate the characteristics of molecular or intermolecular interactions, with a particular focus being placed on the essential method for understanding hydrogen bond properties. Quantification of the topological parameters at the bond critical point (BCP) yields crucial insights into describing potential intermolecular interactions. In this study, we assessed the strength of hydrogen bonds in PRO and its salts using electron density (ρ), Laplacian operator (∇2ρ), electron kinetic density (G), electron potential density (V), and total electron energy density (H) [69]. The AIM topological pathway is illustrated in Figure 12, while the topological parameters of the hydrogen bond BCP are presented in Table 5.
As shown in Figure 12, the BCPs and bond path are marked with orange dots and lines to indicate the complex interactions between the PRO parent crystal and its salt crystals. Additionally, Table 5 indicates that the ranges of ρ and ∇2ρ are 0.0008–0.0311 a.u. and 0.0039–0.1332 a.u., respectively. The values of BCPs are within the acceptable range for hydrogen bonds, signifying the presence of hydrogen bonds in PRO, (PRO)(4AMP)+, and (PRO)(4DAP)+. Furthermore, the nature of hydrogen bonds can be assessed through ∇2ρ and H values. Rozas et al. [70] have established the following standard guidelines for hydrogen bond interactions: when ∇2ρ < 0, it represents a strong covalent interaction; when ∇2ρ > 0 and H < 0, it represents a partially covalent interaction; when ∇2ρ > 0 and H > 0, it represents a weak electrostatic interaction. Theoretical foundations are provided by these standards for distinguishing the strength levels of hydrogen bonds. Emamian and Lu’s method for calculating the hydrogen bonding energy (EH) in a neutral system is shown in Equation (2) [54].
EH = −223.08 × ρ(BCP)/(a.u.) + 0.7423
The calculation method for the hydrogen bond energy (EH) of an electrostatic system such as salt is shown in Equation (3).
EH = −332.34 × ρ(BCP)/(a.u.) − 1.0661
The calculated hydrogen bond energies (EH) of the original PRO crystal form, and its salts are summarized in Table 5.
In the crystal structure of the PRO parent crystal, the primary hydrogen bonds occur between the carboxyl groups of PRO dimers with a strength of −6.79 kcal/mol. Upon salt formation, disruption of the carboxyl dimer in PRO occurs. In (PRO)(4AMP)+, the main hydrogen bonds are formed between the carboxyl group of PRO and the N atom on the pyridine ring of 4AMP through charge-assisted hydrogen bonding (N2+-H2⋯O2), as well as between the carboxyl group of PRO and the amino group of 4AMP through hydrogen bonding (N3-H3A⋯O1 and N3-H3B⋯O2), with strengths of −10.95 kcal/mol, −8.50 kcal/mol, and −7.54 kcal/mol, respectively. In the crystal structure of (PRO)(4DAP)+, the primary hydrogen bond is the charge-assisted hydrogen bond (N2+-H2⋯O1) between the carboxyl group of PRO and the N atom on the pyridine ring of 4DAP, its strength is −11.41 kcal/mol. Computational results indicate that that the main hydrogen bond strength of (PRO)(4DAP)+ is greater than that of (PRO)(4AMP)+, so the melting point is higher; however, the latter exhibits a more robust N-H⋯O hydrogen bond network, resulting in enhanced thermal stability.

3.3.5. IGM Analysis

Molecular cluster fragments containing all types of hydrogen bonds were extracted from PRO and its salts. The intermolecular interactions in PRO and the two salt molecular clusters were analyzed using an IGM. A δginter contour map filled with sign (λ2) ρ colors was generated, as depicted in Figure 13. Different interaction types were distinguished by varying ellipsoid colors: green denoting van der Waals forces and blue representing hydrogen bonds. In the PRO single crystal dimer, green-wrapped blue oblate ellipsoids indicated the presence of van der Waals interactions and hydrogen bonds, with the latter being predominantly present.
After the formation of salt, the carboxyl dipolar groups of PRO are disrupted, leading to the development of a more intricate hydrogen bond network. Additionally, there are predominant blue ellipsoids and smaller green isosurfaces present between molecules in both salts, indicating that van der Waals forces contribute to the stacking of PRO salts alongside hydrogen bonds. The structural analysis of the δginter contour set at 0.01 a.u. reveals that reducing this value results in the emergence of extensive green isosurfaces (Figure S8). This suggests an abundance of weak interactions, such as C-H⋯π and π⋯π, within the PRO salt structure, potentially leading to reduced crystal strength and a lower melting point compared to PRO alone. Furthermore, (PRO)(4AMP)+ exclusively exhibits N2+-H2⋯O1 hydrogen bonds corresponding to blue-dominated ellipsoidal bodies, while, in (PRO)(4AMP)+, there are a greater number of blue-dominated ellipsoidal bodies, indicative of a more robust network of N-H⋯O hydrogen bonds, resulting in enhanced thermal stability.

3.3.6. Lattice Energy (EL) and Hydration-Free Energy (EHF) Analysis

In general, enhanced solubility is typically associated with the increase in the Gibbs free energy of solvation. From a thermodynamic standpoint, the process of hydration can be delineated into sublimation, where individual components of the crystal are transported from an infinite distance into the gas phase and, subsequently, hydration occurs when molecules or ions are enveloped by interacting with water molecules. Consequently, the Gibbs free energy of solvation should equate to both the sublimation free energy and the hydration-free energy (EHF). Lattice energy (EL) denotes the absorbed energy when a crystal transitions to a gaseous state under standard conditions; thus, sublimation free energy can be approximated as negative EL. By computing EL and EHF, we gain deeper insights into PRO’s solubility behavior after salt formation. As shown in Table 6, salt formation disrupts PRO’s original crystal form dimer structure, leading to reduced EL after the salt formation, which is a significant factor contributing to the decrease of melting points. Furthermore, there is consistency between EL order and melting point: (PRO)(4AMP)+ < (PRO)(4DAP)+ < PRO.
Furthermore, the data indicate that the EHF after salt formation of PRO is higher than that of its original crystalline form. However, due to the use of different methods for calculating EL and EHF in this study, the Gibbs free energy of solvation cannot be simply obtained by their summation. Nevertheless, salt formation results in a decrease in EL and an increase in EHF. This leads to a significantly lower Gibbs free energy for the salt solution compared to PRO. Additionally, (PRO)(4AMP)+ exhibits lower EL and higher EHF than (PRO)(4DAP)+, which contributes significantly to its enhanced solubility.

4. Conclusions

This study calculated the FIM of PRO and selected appropriate coformers based on supramolecular synthon rules, subsequently identifying (PRO)(4AMP)+ and (PRO)(4DAP)+ salts through liquid-assisted grinding (LAG). The salts underwent comprehensive characterization using PXRD, DSC, TGA, FTIR, SEM, etc., revealing that PRO displayed various PXRD patterns, melting points, thermal decomposition temperatures, IR absorption peaks, and crystal morphology after salt formation. SCXRD analysis demonstrated distinct hydrogen bond networks, molecular conformations, and packing modes in the two salts. In the phosphate-buffered solution (pH = 6.8), the solubility of the two salts increased to 142.83 and 7.75 times, respectively, compared to that of the raw drug; no crystalline transformation occurred during the 8 weeks of accelerated stability experiments at 40 °C with 75% RH. Molecular intermolecular interactions in the salts’ crystal structure were analyzed using HS; AIM and IGM were employed to analyze the salts’ hydrogen bond networks; MEPs and pKa rules were utilized to investigate salt formation mechanisms and binding sites. Additionally, after salt formation, PRO exhibited decreased lattice energy but increased hydration-free energy, resulting in a lower melting point than that of PRO while demonstrating improved solubility in water. In summary, in this paper concerning the salt formation of soluble drugs, a comprehensive characterization and crystal structure analysis of the prepared PRO salt were conducted. Simultaneously, the principle underlying the changes in the melting point and solubility was deeply explored through quantitative analysis, providing novel insights for enhancing the solubility of drugs by salt formation through the selection of appropriate co-formers. Nevertheless, 4AMP and 4DAP possess certain toxicity, thus more consideration should be given to compounds that comply with GRAS safety certification when choosing cofomers.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14070670/s1, Figure S1: The PXRD patterns of the raw materials and the samples after liquid-assisted grinding experiment; Figure S2: The DSC curves of PRO, 4AMP, 4DAP and PRO-4AMP, PRO-4DAP after the liquid-assisted grinding experiment; Figure S3: The angles between the PRO benzene ring and the coformers’ pyridine ring were measured in both salts. (a) (PRO)(4AMP)+ (b) (PRO)(4DAP)+; Figure S4: FTIR spectra of PRO, 4AMP, 4DAP, (PRO)(4AMP)+, and (PRO)(4DAP)+; Figure S5: HPLC working curve of PRO; Figure S6: PXRD patterns of PRO, (PRO)(4AMP)+, and (PRO)(4DAP)+ before and after the equilibrium solubility experiment (PBS 6.8); Figure S7: PXRD patterns of PRO, (PRO)(4AMP)+, and (PRO)(4DAP)+ before and after the accelerated stability experiment; Figure S8: 0.004 a.u. equivalent surface based on IGM Analysis. (a) (PRO)(4AMP)+ (b) (PRO)(4DAP)+ [55]; Table S1: Compounds used in this experiment; Table S2: C-H···π weak interaction information of (PRO)(4AMP)+ and (PRO)(4DAP)+ [40]; Table S3: π···π weak interaction information of (PRO)(4AMP)+ and (PRO)(4DAP)+ [40]; Table S4: Absorption peaks information of the infrared spectrum of PRO, 4AMP, 4DAP, (PRO)(4AMP)+, and (PRO)(4DAP)+; Table S5: Percentage contribution of various molecular interactions in (PRO)(4AMP)+ and (PRO)(4DAP)+ to the Hirshfeld surface area [44,45]; Table S6: The pKa values of PRO, 4AMP, and 4DAP [60,61].

Author Contributions

Conceptualization, M.Z. and X.H.; methodology, M.Z., X.H. and W.C.; software, M.Z., F.Y. and L.Z. (Liang Zhang); validation, M.Z., X.H. and F.Y.; formal analysis, M.Z.; investigation, M.Z., F.Y. and L.Z. (Liang Zhang); resources, M.Z.; data curation, M.Z.; writing—original draft preparation, M.Z., X.H. and F.Y.; writing—review and editing, M.Z., B.H., L.Z. (Lina Zhou), C.X., S.W. and W.C.; visualization, M.Z., S.W. and W.C.; supervision, W.C.; project administration, S.W. and W.C.; data curation, M.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The CIF files have been deposited at the Cambridge Crystallographic Data Centre (CCDC) with the following numbers: 2283803 and 2360637. These data can be obtained free of charge via https://www.ccdc.cam.ac.uk/structures (accessed on 5 June 2024). The original contributions presented in the study are included in the manuscript and Supplementary Materials; further inquiries can be directed to the corresponding authors.

Acknowledgments

We are grateful for the support and assistance provided by the School of Pharmaceutical Science and Technology at Tianjin University in terms of experimental equipment.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. 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]
  2. Scheepers, M.C.; Lemmerer, A. Exploring the Crystal Structure Landscape of 3,5-Dinitrobenzoic Acid through Various Multicomponent Molecular Complexes. Cryst. Growth Des. 2021, 21, 344–356. [Google Scholar] [CrossRef]
  3. Surov, A.O.; Voronin, A.P.; Vasilev, N.A.; Ilyukhin, A.B.; Perlovich, G.L. Novel cocrystals of the potent 1,2,4-thiadiazole-based neuroprotector with carboxylic acids: Virtual screening, crystal structures and solubility performance. New J. Chem. 2021, 45, 3034–3047. [Google Scholar] [CrossRef]
  4. Babu, N.J.; Nangia, A. Solubility Advantage of Amorphous Drugs and Pharmaceutical Cocrystals. Cryst. Growth Des. 2011, 11, 2662–2679. [Google Scholar] [CrossRef]
  5. Hossain Mithu, M.S.; Economidou, S.; Trivedi, V.; Bhatt, S.; Douroumis, D. Advanced Methodologies for Pharmaceutical Salt Synthesis. Cryst. Growth Des. 2021, 21, 1358–1374. [Google Scholar] [CrossRef]
  6. Zhang, G.; Zhang, L.; Yang, D.; Zhang, N.; He, L.; Du, G.; Lu, Y. Salt screening and characterization of ciprofloxacin. Acta Crystallogr. Sect. B-Struct. Sci. Cryst. Eng. Mat. 2016, 72, 20–28. [Google Scholar] [CrossRef] [PubMed]
  7. Gao, L.; Li, X.; Yan, X.; Zhang, X. Ethylenediamine Salt Enhances the Solubility and Dissolution of Flurbiprofen. ChemistryOpen 2024, 13, e202300262. [Google Scholar] [CrossRef]
  8. Li, C.; Wu, D.; Gao, Z.; Chen, W. Molecular simulation studies on the design of multicomponent sorbic acid crystals for tuning solubility. CrystEngComm 2023, 25, 4889–4901. [Google Scholar] [CrossRef]
  9. Li, J.; Wu, D.; Xiao, Y.; Li, C.; Ji, X.; Sun, Q.; Chang, D.; Zhou, L.; Jing, D.; Gong, J.; et al. Salts of 2-hydroxybenzylamine with improvements on solubility and stability: Virtual and experimental screening. Eur. J. Pharm. Sci. 2022, 169, 106091. [Google Scholar] [CrossRef]
  10. Sun, J.; Jia, L.; Wang, M.; Liu, Y.; Li, M.; Han, D.; Gong, J. Novel Drug–Drug Multicomponent Crystals of Epalrestat–Metformin: Improved Solubility and Photostability of Epalrestat and Reduced Hygroscopicity of Metformin. Cryst. Growth Des. 2022, 22, 1005–1016. [Google Scholar] [CrossRef]
  11. Hu, X.; Xiao, Y.; Qi, L.; Bai, Y.; Sun, Y.; Ye, Y.; Xie, C. Enhancing the dissolution and bacteriostatic activity of trimethoprim through salt formation. CrystEngComm 2023, 25, 3445–3459. [Google Scholar] [CrossRef]
  12. Kumar, L.; Amin, A.; Bansal, A.K. An overview of automated systems relevant in pharmaceutical salt screening. Drug Discov. Today 2007, 12, 1046–1053. [Google Scholar] [CrossRef] [PubMed]
  13. Perez-Ruiz, F.; Herrero-Beites, A.M.; Atxotegi Saenz De Buruaga, J. Chapter 12—Uricosuric Therapy of Hyperuricemia in Gout; W.B. Saunders: Philadelphia, PA, USA, 2012. [Google Scholar]
  14. Sattui, S.E.; Gaffo, A.L. Treatment of hyperuricemia in gout: Current therapeutic options, latest developments and clinical implications. Ther. Adv. Musculoskelet. Dis. 2016, 8, 145–159. [Google Scholar] [CrossRef] [PubMed]
  15. Barone, S.; Xu, J.; Zahedi, K.; Brooks, M.; Soleimani, M. Probenecid Downregulates Kidney Pendrin and AQP-2 and Potentiates Hydrochlorothiazide Diuresis. Front. Physiol. 2018, 9, A325. [Google Scholar] [CrossRef] [PubMed]
  16. Wei, R.; Wang, J.; Xu, Y.; Yin, B.; He, F.; Du, Y.; Peng, G.; Luo, B. Probenecid protects against cerebral ischemia/reperfusion injury by inhibiting lysosomal and inflammatory damage in rats. Neuroscience 2015, 301, 168–177. [Google Scholar] [CrossRef] [PubMed]
  17. Smetanová, L.; Stĕtinová, V.; Svoboda, Z.; Kvetina, J. Caco-2 cells, biopharmaceutics classification system (BCS) and biowaiver. Acta Medica 2011, 54, 3–8. [Google Scholar] [PubMed]
  18. Gisclon, L.G.; Boyd, R.A.; Williams, R.L.; Giacomini, K.M. The effect of probenecid on the renal elimination of cimetidine. Clin. Pharmacol. Ther. 1989, 45, 444–452. [Google Scholar] [CrossRef]
  19. Bruni, G.; Monteforte, F.; Maggi, L.; Friuli, V.; Ferrara, C.; Mustarelli, P.; Girella, A.; Berbenni, V.; Capsoni, D.; Milanese, C.; et al. Probenecid and benzamide: Cocrystal prepared by a green method and its physico-chemical and pharmaceutical characterization. J. Therm. Anal. Calorim. 2020, 140, 1859–1869. [Google Scholar] [CrossRef]
  20. Shen, Y.; Zong, S.; Dang, L.; Wei, H. Solubility and thermodynamics of probenecid-4,4′-azopyridine cocrystal in pure and binary solvents. J. Mol. Liq. 2019, 290, 111195. [Google Scholar] [CrossRef]
  21. Gupta, P.; Karothu, D.P.; Ahmed, E.; Naumov, P.; Nath, N.K. Thermally Twistable, Photobendable, Elastically Deformable, and Self-Healable Soft Crystals. Angew. Chem. Int. Ed. 2018, 57, 8498–8502. [Google Scholar] [CrossRef]
  22. Nath, N.K.; Hazarika, M.; Gupta, P.; Ray, N.R.; Paul, A.K.; Nauha, E. Plastically bendable crystals of probenecid and its cocrystal with 4,4′-Bipyridine. J. Mol. Struct. 2018, 1160, 20–25. [Google Scholar] [CrossRef]
  23. Rao, K.U.; Bhogala, B.R.; Maguire, A.R.; Lawrence, S.E. Symmetry assisted tuning of bending and brittle multi-component forms of probenecid. Chem. Commun. 2017, 53, 3381–3384. [Google Scholar]
  24. Bruni, G.; Monteforte, F.; Maggi, L.; Girella, A.; Berbenni, V.; Milanese, C.; Marini, A. Probenecid and benzamide: DSC applied to the study of an “impossible” pharmaceutical system. J. Therm. Anal. Calorim. 2021, 145, 391–402. [Google Scholar] [CrossRef]
  25. Li, Y.; Zhang, Y.; Xu, H.; Li, M.; Li, Z.; Song, Z.; Chen, J.; Liu, Y.; Sun, Y.; Yang, Z. Synthesis and characterization of supramolecular assembly probenecid cocrystal. J. Mol. Struct. 2024, 1298, 136786. [Google Scholar] [CrossRef]
  26. Haneef, J.; Markad, D.; Chadha, R. Interaction Map Driven Cocrystallization of Ambrisentan: Structural and Biopharmaceutical Evaluation. Cryst. Growth Des. 2020, 20, 4612–4620. [Google Scholar] [CrossRef]
  27. Solares-Briones, M.; Coyote-Dotor, G.; Páez-Franco, J.C.; Zermeño-Ortega, M.R.; de la O Contreras, C.M.; Canseco-González, D.; Avila-Sorrosa, A.; Morales-Morales, D.; Germán-Acacio, J.M. Mechanochemistry: A Green Approach in the Preparation of Pharmaceutical Cocrystals. Pharmaceutics 2021, 13, 790. [Google Scholar] [CrossRef]
  28. Xiao, Y.; Wu, C.; Hu, X.; Chen, K.; Qi, L.; Cui, P.; Zhou, L.; Yin, Q. Mechanochemical Synthesis of Cocrystal: From Mechanism to Application. Cryst. Growth Des. 2023, 23, 4680–4700. [Google Scholar] [CrossRef]
  29. Khan, E.; Shukla, A.; Srivastava, K.; Gangopadhyay, D.; Assi, K.H.; Tandon, P.; Vangala, V.R. Structural and Reactivity Analyses of Nitrofurantoin–4-dimethylaminopyridine Salt Using Spectroscopic and Density Functional Theory Calculations. Crystals 2019, 9, 413. [Google Scholar] [CrossRef]
  30. Chitra, R.; Choudhury, R.R.; Capet, F.; Roussel, P.; Bhatt, P. Crystal structure of 4-aminopyridinium 3-(4-aminopyridinium) succinate tetra hydrate: A new salt from 4-aminopyridine and maleic acid crystallization. J. Mol. Struct. 2021, 1234, 130142. [Google Scholar] [CrossRef]
  31. Boycov, D.E.; Drozd, K.V.; Manin, A.N.; Churakov, A.V.; Perlovich, G.L. New Solid Forms of Nitrofurantoin and 4-Aminopyridine Salt: Influence of Salt Hydration Level on Crystal Packing and Physicochemical Properties. Molecules 2022, 27, 8990. [Google Scholar] [CrossRef]
  32. Huang, S.; Venables, D.S.; Lawrence, S.E. Pharmaceutical Salts of Piroxicam and Meloxicam with Organic Counterions. Cryst. Growth Des. 2022, 22, 6504–6520. [Google Scholar] [CrossRef]
  33. Ma, D.J.; Pei, T.Z.; Bai, Y.H.; Zhou, L.N.; Bao, Y.; Yin, Q.X.; Xie, C. Salts formation between ibuprofen and pyridine derivatives: Effect of amino group on supramolecular packing and proton transfer. J. Mol. Struct. 2019, 1179, 487–494. [Google Scholar] [CrossRef]
  34. Nauha, E.; Bernstein, J. “Predicting” Polymorphs of Pharmaceuticals Using Hydrogen Bond Propensities: Probenecid and its Two Single-Crystal-to-Single-Crystal Phase Transitions. J. Pharm. Sci. 2015, 104, 2056–2061. [Google Scholar] [CrossRef]
  35. Desiraju, G.R. The Supramolecular Synthon in Crystal Engineering. In Stimulating Concepts in Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2000. [Google Scholar]
  36. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A 2008, 64, 112–122. [Google Scholar] [CrossRef]
  37. Corpinot, M.K.; Bučar, D. A Practical Guide to the Design of Molecular Crystals. Cryst. Growth Des. 2019, 19, 1426–1453. [Google Scholar] [CrossRef]
  38. Devogelaer, J.; Brugman, S.J.T.; Meekes, H.; Tinnemans, P.; Vlieg, E.; de Gelder, R. Cocrystal design by network-based link prediction. CrystEngComm 2019, 21, 6875–6885. [Google Scholar] [CrossRef]
  39. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Crystallogr. 2015, 48, 3–10. [Google Scholar] [CrossRef] [PubMed]
  40. Spek, A.L. Single-crystal structure validation with the program PLATON. J. Appl. Crystallogr. 2003, 36, 7–13. [Google Scholar] [CrossRef]
  41. Macrae, C.F.; Edgington, P.R.; Mccabe, P.; Pidcock, E.; Shields, G.P.; Taylor, R.; Towler, M.; van de Streek, J. Mercury: Visualization and analysis of crystal structures. J. Appl. Crystallogr. 2006, 39, 453–457. [Google Scholar] [CrossRef]
  42. Zhou, L.; Yang, L.; Tilton, S.; Wang, J. Development of a high throughput equilibrium solubility assay using miniaturized shake-flask method in early drug discovery. J. Pharm. Sci. 2007, 96, 3052–3071. [Google Scholar] [CrossRef]
  43. Yang, R. Improvement of the determination method for the content of probenecid. Chin. Med. Dig. Intern. Med. 2012, 7, 408–409. [Google Scholar]
  44. Mckinnon, J.J.; Spackman, M.A.; Mitchell, A.S. Novel tools for visualizing and exploring intermolecular interactions in molecular crystals. Acta Crystallogr. Sect. B 2004, 60, 627–668. [Google Scholar] [CrossRef]
  45. Spackman, P.R.; Turner, M.J.; Mckinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer: A program for Hirshfeld surface analysis, visualization and quantitative analysis of molecular crystals. J. Appl. Crystallogr. 2021, 54, 1006–1011. [Google Scholar] [CrossRef]
  46. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Kristallographie—Cryst. Mater. 2005, 220, 567–570. [Google Scholar] [CrossRef]
  47. Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar]
  48. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892–7895. [Google Scholar] [CrossRef] [PubMed]
  49. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  50. Eichkorn, K.; Treutler, O.; Öhm, H.; Häser, M.; Ahlrichs, R. Auxiliary basis sets to approximate Coulomb potentials. Chem. Phys. Lett. 1995, 240, 283–290. [Google Scholar] [CrossRef]
  51. Mardirossian, N.; Head-Gordon, M. ΩB97M-V: A combinatorially optimized, range-separated hybrid, meta-GGA density functional with VV10 nonlocal correlation. J. Chem. Phys. 2016, 144, 214110. [Google Scholar] [CrossRef]
  52. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297. [Google Scholar] [CrossRef]
  53. Neese, F.; Wennmohs, F.; Becker, U.; Riplinger, C. The ORCA quantum chemistry program package. J. Chem. Phys. 2020, 152, 224108. [Google Scholar] [CrossRef] [PubMed]
  54. Emamian, S.; Lu, T.; Kruse, H.; Emamian, H. Exploring Nature and Predicting Strength of Hydrogen Bonds: A Correlation Analysis Between Atoms-in-Molecules Descriptors, Binding Energies, and Energy Components of Symmetry-Adapted Perturbation Theory. J. Comput. Chem. 2019, 40, 2868–2881. [Google Scholar] [CrossRef] [PubMed]
  55. Lu, T.; Chen, Q. Independent gradient model based on Hirshfeld partition: A new method for visual study of interactions in chemical systems. J. Comput. Chem. 2022, 43, 539–555. [Google Scholar] [CrossRef] [PubMed]
  56. Zhang, J.; Lu, T. Efficient evaluation of electrostatic potential with computerized optimized code. Phys. Chem. Chem. Phys. 2021, 23, 20323–20328. [Google Scholar] [CrossRef]
  57. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  58. Lu, T.; Chen, F. Quantitative analysis of molecular surface based on improved Marching Tetrahedra algorithm. J. Mol. Graph. Model. 2012, 38, 314–323. [Google Scholar] [CrossRef]
  59. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  60. Mittapalli, S.; Mannava, M.K.C.; Sahoo, R.; Nangia, A. Cocrystals, Salts, and Supramolecular Gels of Nonsteroidal Anti-Inflammatory Drug Niflumic Acid. Cryst. Growth Des. 2019, 19, 219–230. [Google Scholar] [CrossRef]
  61. Dzotam, J.K.; Simo, I.K.; Bitchagno, G.; Celik, I.; Sandjo, L.P.; Tane, P.; Kuete, V. In vitro antibacterial and antibiotic modifying activity of crude extract, fractions and 3’,4’,7-trihydroxyflavone from Myristica fragrans Houtt against MDR Gram-negative enteric bacteria. BMC Complement. Altern. Med. 2018, 18, 15. [Google Scholar] [CrossRef]
  62. Vener, M.V.; Levina, E.O.; Koloskov, O.A.; Rykounov, A.A.; Voronin, A.P.; Tsirelson, V.G. Evaluation of the Lattice Energy of the Two-Component Molecular Crystals Using Solid-State Density Functional Theory. Cryst. Growth Des. 2014, 14, 4997–5003. [Google Scholar] [CrossRef]
  63. Ranke, J.; Othman, A.; Fan, P.; Müller, A. Explaining Ionic Liquid Water Solubility in Terms of Cation and Anion Hydrophobicity. Int. J. Mol. Sci. 2009, 10, 1271–1289. [Google Scholar] [CrossRef]
  64. Vyboishchikov, S.F.; Voityuk, A.A. Fast non-iterative calculation of solvation energies for water and non-aqueous solvents. J. Comput. Chem. 2021, 42, 1184–1194. [Google Scholar] [CrossRef] [PubMed]
  65. Swapna, B.; Nangia, A. Epalrestat–Cytosine Cocrystal and Salt Structures: Attempt to Control E,Z → Z,Z Isomerization. Cryst. Growth Des. 2017, 17, 3350–3360. [Google Scholar] [CrossRef]
  66. Janiak, C. A critical account on π–π stacking in metal complexes with aromatic nitrogen-containing ligands. J. Chem. Soc. Dalton Trans. 2000, 21, 3885–3896. [Google Scholar] [CrossRef]
  67. Cruz-Cabeza, A.J. Acid-base crystalline complexes and the pKa rule. CrystEngComm 2012, 14, 6362–6365. [Google Scholar] [CrossRef]
  68. Tothadi, S.; Shaikh, T.R.; Gupta, S.; Dandela, R.; Vinod, C.P.; Nangia, A.K. Can we Identify the Salt–Cocrystal Continuum State Using XPS? Cryst. Growth Des. 2021, 21, 735–747. [Google Scholar] [CrossRef]
  69. Espinosa, E.; Molins, E.; Lecomte, C. Hydrogen bond strengths revealed by topological analyses of experimentally observed electron densities. Chem. Phys. Lett. 1998, 285, 170–173. [Google Scholar] [CrossRef]
  70. Rozas, I.; Alkorta, I.; Elguero, J. Behavior of Ylides Containing N, O, and C Atoms as Hydrogen Bond Acceptors. J. Am. Chem. Soc. 2000, 122, 11154–11161. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of (a) PRO, (b) 4AMP, (c) 4DAP, (d) (PRO)(4AMP)+ and (e) (PRO)(4DAP)+.
Figure 1. Chemical structures of (a) PRO, (b) 4AMP, (c) 4DAP, (d) (PRO)(4AMP)+ and (e) (PRO)(4DAP)+.
Crystals 14 00670 g001
Figure 2. (a) The 3D full interaction map of PRO, with the possible regions of acceptors shown in blue and the donors in red. (b) Potential sites of molecular interactions of PRO marked with orange dashed boxes. (c) Supramolecular synthons of PRO salts.
Figure 2. (a) The 3D full interaction map of PRO, with the possible regions of acceptors shown in blue and the donors in red. (b) Potential sites of molecular interactions of PRO marked with orange dashed boxes. (c) Supramolecular synthons of PRO salts.
Crystals 14 00670 g002
Figure 3. (a) Asymmetric unit of (PRO)(4AMP)+. (b) 1D chain of (PRO)(4AMP)+ viewed along the b-axis. (c) 1D double chain of (PRO)(4AMP)+ viewed along the b-axis (yellow and green represent different 1D chains). (d) 3D packing diagram of (PRO)(4AMP)+ viewed along the a-axis (where red and blue represent (PRO) anion and (4DAP)+ cation, respectively).
Figure 3. (a) Asymmetric unit of (PRO)(4AMP)+. (b) 1D chain of (PRO)(4AMP)+ viewed along the b-axis. (c) 1D double chain of (PRO)(4AMP)+ viewed along the b-axis (yellow and green represent different 1D chains). (d) 3D packing diagram of (PRO)(4AMP)+ viewed along the a-axis (where red and blue represent (PRO) anion and (4DAP)+ cation, respectively).
Crystals 14 00670 g003
Figure 4. (a) Asymmetric unit of (PRO)(4DAP)+. (b) 1D chain of the (PRO)(4DAP)+ salt observed along the b-axis. (c) 2D planar structure of the (PRO)(4DAP)+ salt observed along the a-axis (brown, green, and purple representing distinct 1D chains). (d) 3D packing structure of the (PRO)(4DAP)+ salt observed along the a-axis (where red and blue represent the (PRO) anion and (4DAP)+ cation, respectively, while green and yellow represent opposing direction 2D bilayers).
Figure 4. (a) Asymmetric unit of (PRO)(4DAP)+. (b) 1D chain of the (PRO)(4DAP)+ salt observed along the b-axis. (c) 2D planar structure of the (PRO)(4DAP)+ salt observed along the a-axis (brown, green, and purple representing distinct 1D chains). (d) 3D packing structure of the (PRO)(4DAP)+ salt observed along the a-axis (where red and blue represent the (PRO) anion and (4DAP)+ cation, respectively, while green and yellow represent opposing direction 2D bilayers).
Crystals 14 00670 g004
Figure 5. (a) Superposition diagram of different conformations in PRO parent crystal and PRO salt single crystal (red, orange and purple represent PRO molecules in PRO, (PRO)(4AMP)+ and (PRO)(4DAP)+, respectively). (b) Comparison of the two salt crystal accumulations of PRO (where the green molecule represents (PRO)(4DAP)+, the blue circle represents PRO molecules at the same stacking position). (c) Hydrogen bond networks of (PRO)(4AMP)+. (d) Hydrogen bond networks of (PRO)(4DAP)+.
Figure 5. (a) Superposition diagram of different conformations in PRO parent crystal and PRO salt single crystal (red, orange and purple represent PRO molecules in PRO, (PRO)(4AMP)+ and (PRO)(4DAP)+, respectively). (b) Comparison of the two salt crystal accumulations of PRO (where the green molecule represents (PRO)(4DAP)+, the blue circle represents PRO molecules at the same stacking position). (c) Hydrogen bond networks of (PRO)(4AMP)+. (d) Hydrogen bond networks of (PRO)(4DAP)+.
Crystals 14 00670 g005
Figure 6. PXRD patterns of PRO, 4AMP, 4DAP, experimental and calculated (PRO)(4AMP)+ and (PRO)(4DAP)+.
Figure 6. PXRD patterns of PRO, 4AMP, 4DAP, experimental and calculated (PRO)(4AMP)+ and (PRO)(4DAP)+.
Crystals 14 00670 g006
Figure 7. DSC curves of PRO, 4AMP, 4DAP, (PRO)(4AMP)+ and (PRO)(4DAP)+.
Figure 7. DSC curves of PRO, 4AMP, 4DAP, (PRO)(4AMP)+ and (PRO)(4DAP)+.
Crystals 14 00670 g007
Figure 8. TGA-DSC curves of (PRO)(4AMP)+ and (PRO)(4DAP)+ (Solid lines represent the TGA curves, dashed lines represent the DSC curves).
Figure 8. TGA-DSC curves of (PRO)(4AMP)+ and (PRO)(4DAP)+ (Solid lines represent the TGA curves, dashed lines represent the DSC curves).
Crystals 14 00670 g008
Figure 9. SEM spectra of (a) PRO, (b) 4AMP, (c) 4DAP, (d) (PRO)(4AMP)+ and (e) (PRO)(4DAP)+.
Figure 9. SEM spectra of (a) PRO, (b) 4AMP, (c) 4DAP, (d) (PRO)(4AMP)+ and (e) (PRO)(4DAP)+.
Crystals 14 00670 g009
Figure 10. 3D Hirshfeld surface maps of (a) (PRO)(4AMP)+ and (b) (PRO)(4DAP)+ (dnorm, dshape, and dcurve were mapped, respectively). 2D fingerprint plots of (c) (PRO)(4AMP)+ and (d) (PRO)(4DAP)+.
Figure 10. 3D Hirshfeld surface maps of (a) (PRO)(4AMP)+ and (b) (PRO)(4DAP)+ (dnorm, dshape, and dcurve were mapped, respectively). 2D fingerprint plots of (c) (PRO)(4AMP)+ and (d) (PRO)(4DAP)+.
Crystals 14 00670 g010
Figure 11. MEP mapped on the 0.001 a.u. electron density isosurface of (a) PRO, (b) 4AMP, (c) 4DAP, (d) Asymmetric unit of (PRO)(4AMP)+, and (e) Asymmetric unit of (PRO)(4DAP)+. The local maximum and minimum values (kcal/mol) are marked with red and blue texts, respectively.
Figure 11. MEP mapped on the 0.001 a.u. electron density isosurface of (a) PRO, (b) 4AMP, (c) 4DAP, (d) Asymmetric unit of (PRO)(4AMP)+, and (e) Asymmetric unit of (PRO)(4DAP)+. The local maximum and minimum values (kcal/mol) are marked with red and blue texts, respectively.
Crystals 14 00670 g011
Figure 12. Topological geometric structures of PRO, (PRO)(4AMP)+ and (PRO)(4DAP)+, with purple, orange, yellow, and green spheres corresponding to the (3, −3), (3, −1), (3, +1), and (3, +3) critical points of electron density (nuclear, bond, ring, and cage critical points). The orange curves correspond to the bond path. (a) PRO. (b) (PRO)(4AMP)+. (c) (PRO)(4DAP)+.
Figure 12. Topological geometric structures of PRO, (PRO)(4AMP)+ and (PRO)(4DAP)+, with purple, orange, yellow, and green spheres corresponding to the (3, −3), (3, −1), (3, +1), and (3, +3) critical points of electron density (nuclear, bond, ring, and cage critical points). The orange curves correspond to the bond path. (a) PRO. (b) (PRO)(4AMP)+. (c) (PRO)(4DAP)+.
Crystals 14 00670 g012
Figure 13. 0.01 a.u. equivalent surface based on IGM Analysis, (a) PRO, (b) (PRO)(4AMP)+ and (c) (PRO)(4DAP)+ (Green represent van der Waals forces, blue represent hydrogen bonds).
Figure 13. 0.01 a.u. equivalent surface based on IGM Analysis, (a) PRO, (b) (PRO)(4AMP)+ and (c) (PRO)(4DAP)+ (Green represent van der Waals forces, blue represent hydrogen bonds).
Crystals 14 00670 g013
Table 1. Crystallographic data and structural refinement details of (PRO)(4AMP)+ and (PRO)(4DAP)+.
Table 1. Crystallographic data and structural refinement details of (PRO)(4AMP)+ and (PRO)(4DAP)+.
Compound(PRO)(4AMP)+(PRO)(4DAP)+
Empirical formulaC18H25N3O4SC20H29N3O4S
Formula weight379.47407.52
Temperature/K293 (2)293 (2)
Crystal systemorthorhombicmonoclinic
Space groupP212121P21/c
a/Å7.1146 (1)6.8830 (1)
b/Å11.2314 (2)7.1512 (2)
c/Å24.3877 (4)43.1607 (8)
α/°9090
β/°9093.680 (2)
γ/°9090
Volume/Å31948.75 (5)2120.06 (8)
Z44
ρcalc/(g/cm3)1.2931.277
F (000)808.0872.0
2θ range for data collection/°3.340 to 72.2586.002 to 57.394
Goodness-of-fit on F21.0251.100
Rint0.06130.0502
R1 indexes [I ≥ 2σ (I)]0.05900.0534
ωR2 indexes [all data]0.12800.1496
CCDC No.22838032360637
Table 2. C-O bond length distribution of the salts of PRO.
Table 2. C-O bond length distribution of the salts of PRO.
CompoundDC-O (Å)ΔDC-O (Å)Proton Transfer
(PRO)(4AMP)+C1-O1:1.251, C1-O2:1.2630.012Yes
(PRO)(4DAP)+C1-O1:1.265, C1-O2:1.2410.024Yes
Table 3. Hydrogen bonding parameters of (PRO)(4AMP)+ and (PRO)(4DAP)+.
Table 3. Hydrogen bonding parameters of (PRO)(4AMP)+ and (PRO)(4DAP)+.
D-H⋯AD-H(Å)H⋯A(Å)D⋯A(Å)D-H⋯A(°)Symmetry
(PRO)(4AMP)+
N2+-H2⋯O20.8611.8852.732 (3)167.36
N3-H3A⋯O10.8601.9652.812 (3)167.77x, −1/2 + y, 3/2 − z
N3-H3B⋯O20.8612.0422.888 (3)167.511 − x, −1/2 + y, 3/2 − z
C12-H12B⋯O30.9702.5853.268 (3)127.551 + x, y, z
C14-H14⋯O40.9302.5293.203 (3)129.62−1/2 + x, 1/2 − y, 1 − z
C18-H18⋯O40.9302.5563.357 (3)144.501/2 + x, 1/2 − y, 1 − z
(PRO)(4DAP)+
N2+-H2⋯O10.8591.8582.7130 (19)172.85
C10-H10C⋯O40.9612.5733.465 (19)154.49x, 1 + y, z
C14-H14⋯O20.9302.3503.213 (2)154.29−1 + x, y, z
C18-H18⋯O20.9302.3563.036 (2)129.74
Table 4. (PRO)(4AMP)+ and (PRO)(4DAP)+ equilibrium solubility data (PBS 6.8).
Table 4. (PRO)(4AMP)+ and (PRO)(4DAP)+ equilibrium solubility data (PBS 6.8).
CompoundEquilibrium Solubility (mg/mL)The Increasing Factor
PRO4.61
(PRO)(4AMP)+658.45142.83
(PRO)(4DAP)+35.717.75
Table 5. Hydrogen bond topological parameters for PRO, (PRO)(4AMP)+ and (PRO)(4DAP)+ in AIM analysis.
Table 5. Hydrogen bond topological parameters for PRO, (PRO)(4AMP)+ and (PRO)(4DAP)+ in AIM analysis.
D-H⋯AH⋯A (Å)ρ (a.u.)2ρ (a.u.)G (a.u.)V (a.u.)H (a.u.)EH (kcal/mol)
PRO
O1-H1⋯O21.7960.03380.14340.0343−0.03280.0016−6.79
(PRO)(4AMP)+
N2+-H2⋯O21.8850.02970.12670.0288−0.02600.0028−10.95
N3-H3A⋯O11.9650.02240.10780.0228−0.01860.0042−8.50
N3-H3B⋯O22.0420.01950.08500.0180−0.01470.0033−7.54
C12-H12B⋯O32.5850.00760.02730.0059−0.00490.0010−3.59
C14-H14⋯O42.5290.00800.03080.0064−0.00520.0013−3.71
C18-H18⋯O42.5560.00800.02710.0059−0.00490.0009−3.74
(PRO)(4DAP)+
N2+-H2⋯O11.8580.03110.13320.0305−0.02770.0028−11.41
C10-H10C⋯O42.5730.00080.00390.0007−0.00040.0003−1.33
C14-H14⋯O22.3500.01040.03970.0081−0.00620.0019−4.52
C18-H18⋯O22.3560.01280.04730.0100−0.00820.0018−5.30
Table 6. Calculated values of EL and EHF.
Table 6. Calculated values of EL and EHF.
CompoundEL (kcal/mol)EHF (kcal/mol)
PRO−64.87−14.33
(PRO)(4AMP)+−41.84−96.95
(PRO)(4DAP)+−50.62−95.84
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

Zhang, M.; Hou, X.; Yu, F.; Zhang, L.; Hou, B.; Zhou, L.; Xie, C.; Wu, S.; Chen, W. Synthesis, Characterization, and Analysis of Probenecid and Pyridine Compound Salts. Crystals 2024, 14, 670. https://doi.org/10.3390/cryst14070670

AMA Style

Zhang M, Hou X, Yu F, Zhang L, Hou B, Zhou L, Xie C, Wu S, Chen W. Synthesis, Characterization, and Analysis of Probenecid and Pyridine Compound Salts. Crystals. 2024; 14(7):670. https://doi.org/10.3390/cryst14070670

Chicago/Turabian Style

Zhang, Menglong, Xinyu Hou, Fuhai Yu, Liang Zhang, Baohong Hou, Lina Zhou, Chuang Xie, Songgu Wu, and Wei Chen. 2024. "Synthesis, Characterization, and Analysis of Probenecid and Pyridine Compound Salts" Crystals 14, no. 7: 670. https://doi.org/10.3390/cryst14070670

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