This comprehensive review explores the synthesis, functionalization, and burgeoning biomedical applications of inorganic cluster supramolecular compounds (ICSs).
This comprehensive review explores the synthesis, functionalization, and burgeoning biomedical applications of inorganic cluster supramolecular compounds (ICSs). Targeting researchers, scientists, and drug development professionals, it delves into the foundational design principles of these architectures, including metallacages and coordination complexes, established through coordination chemistry and self-assembly. The article details advanced synthetic methodologies and characterization techniques, highlighting their application in anticancer drug development, targeted drug delivery, bioimaging, and catalysis. It further addresses critical troubleshooting and optimization strategies for stability and reproducibility. Finally, a comparative analysis validates ICSs against traditional materials, evaluating their performance, biocompatibility, and clinical potential to inform future therapeutic innovation.
Inorganic Cluster Supramolecular Compounds (ICSs) represent a advanced class of structures formed via coordination-driven self-assembly between metal-containing nodes and organic ligands. These compounds, also broadly classified as Supramolecular Coordination Complexes (SCCs), occupy a strategic position in modern chemistry, bridging the gap between classical supramolecular chemistry and coordination chemistry [1] [2]. The field has evolved significantly since the early discoveries of crown ethers by Pedersen, cryptands by Lehn, and spherands by Cram, which established the fundamental principles of molecular recognition through non-covalent interactions [3] [2]. The integration of metal-ligand coordination bonds, with energies intermediate (15-50 kcal/mol) between covalent bonds (60-120 kcal/mol) and weak non-covalent interactions (0.5-10 kcal/mol), provides the unique combination of structural stability and dynamic reversibility that characterizes ICSs [2]. This dynamic reversibility allows for self-correction during assembly, leading to discrete, thermodynamically stable architectures with precise spatial control [2].
The significance of ICSs extends across multiple disciplines, from catalysis and sensing to materials science and biomedicine [1] [4]. Their structural precision, tunable cavities, and capacity for molecular recognition make them particularly valuable for pharmaceutical applications, where they can function as novel anticancer agents, drug delivery vehicles, theranostic agents, and biosensors [1] [4] [5]. The rational design of these structures leverages the predictable nature of metal-ligand coordination spheres and the geometric preferences of metal ions to create finite two-dimensional (2D) and three-dimensional (3D) architectures with defined shapes and sizes [2].
ICSs can be systematically classified based on their dimensionality and structural features. The primary categories include metallacages (3D structures), helicates (chiral helical structures), and metallacycles (2D cyclic structures).
Metallacages are three-dimensional supramolecular structures that feature an internal cavity accessible to guest encapsulation [1] [6]. These structures belong to the broader class of metallocavitands and exploit host-guest chemistry for various functions and applications [1]. The internal cavity volume can range significantly, with some structures demonstrating flexibility to adapt from 69 à ³ to 87 à ³ to accommodate different guest molecules [7]. This adaptability enables metallacages to function as "molecular flasks" that can isolate and stabilize reactive intermediates or facilitate unusual chemical transformations [2]. The host-guest chemistry of 3D self-assembled structures within the metallacages family can be exploited to design novel drug delivery systems for anticancer chemotherapeutics [1] [6].
Helicates are chiral supramolecular structures typically formed by two or more organic strands wrapping around metal centers in a helical fashion [8]. The helical structure is governed by the configuration of organic strands at the metal centers, often resulting from interactions between planar chiral building blocks [8]. These structures have been studied extensively for their molecular recognition properties of nucleic acid structures, with possible applications in therapy [1]. Helicates can undergo stereochemical inversion through various mechanisms, such as the Bailar twist, transitioning between enantiomers (ÎÎâÎÎ) via mesocate intermediates [7]. Recent advancements include the construction of double helicates using planar chiral [2.2]paracyclophane (PCP) derivatives, where the helical structure is dominated by the 3D planar chiral units rather than the metal-coordination centers [8].
Metallacycles are two-dimensional cyclic structures formed through coordination-driven self-assembly [2]. These structures include various polygons such as squares, rectangles, triangles, and hexagons, with sizes scalable to nanoscopic dimensions [9] [2]. The Fujita group reported 'Pd48L96', the largest discrete self-assembled edge-directed polyhedron, demonstrating the remarkable scalability of these structures [1] [6]. Metallacycles can be designed to incorporate photochromic units like diarylethene, enabling light-triggered reversible structural transformations with potential applications in molecular switches, smart soft materials, and photodynamic therapy [9].
Table 1: Key Structural Features of Major ICS Classes
| ICS Class | Dimensionality | Key Characteristics | Representative Examples |
|---|---|---|---|
| Metallacages | 3D | Internal cavity for guest encapsulation, structural flexibility | [Pd6L8] cages, Pillarplexes |
| Helicates | 1D/2D/3D | Chiral helical structures, stereochemical inversion | Double helicates, [Ga2(3)3] helicates |
| Metallacycles | 2D | Cyclic polygons, photoresponsive properties | Molecular squares, triangles, hexagons |
The construction of ICSs follows several well-established design principles that leverage the directional nature of metal-ligand coordination bonds.
The directional bonding approach represents a fundamental strategy for constructing ICSs, relying on structurally rigid complementary precursor units with predefined bite angles [2]. This method involves combining donor building blocks (organic ligands with specific angular orientations) and acceptor units (metal-containing subunits with fixed coordination angles) in precise stoichiometric ratios [2]. The symmetry and number of binding sites within each precursor unit dictate the final architecture's shape. For instance, molecular squares can be designed through multiple pathways: combining four 90° angular units with four 180° linear units, or a 2:2 assembly of two different 90° angular units [2]. Similarly, 3D polyhedral architectures require combinations of angular and linear subunits with multidentate precursors featuring more than two binding sites [2].
The symmetry interaction approach, pioneered by Raymond and others, utilizes the inherent symmetry of building blocks to direct the self-assembly process [2]. This method considers the point group symmetry of both organic ligands and metal complexes to predict the resulting supramolecular architecture. The molecular paneling approach, developed by Fujita and referred to as the "paneling method," employs planar multidentate ligand molecules that form the faces of supramolecular polyhedra upon coordination to convergently oriented metal nodes [1] [6]. This face-directed method contrasts with edge-directed approaches that use banana-shaped ligands to form the edges of SCCs [1].
Nitschke introduced the concept of "subcomponent self-assembly," where the actual linker forms in situ through reactions such as imine formation from aldehydes and amines [1] [6]. This approach allows for covalent post-assembly modifications of the SCCs, adding another layer of functionality and complexity [1]. The dynamic covalent chemistry involved in subcomponent self-assembly enables error-checking and self-correction during the formation process, leading to highly defined structures.
Table 2: Quantitative Data on Selected ICS Structures
| ICS Structure | Cavity Volume (à ³) | Photoconversion Yield (%) | Energy Transfer Efficiency (%) | Cytotoxicity (IC50) |
|---|---|---|---|---|
| Cu2(1o)2 Metallacycle [9] | N/A | ~90 | N/A | N/A |
| Diarylethene Hexagon 6 [9] | N/A | ~100 (vs. 88 for ligand) | N/A | N/A |
| [Fe4(6)6]8+ Tetrahedron [7] | 69-87 (flexible) | N/A | N/A | N/A |
| PCP-TPy1/2 Double Helicates [8] | N/A | N/A | Up to 89.3 | N/A |
The synthesis of ICSs typically follows a straightforward procedure involving the combination of metal acceptor and organic donor building blocks in appropriate solvents [9] [8]. A representative protocol for constructing double helicates, as described for PCP-TPy systems, involves stirring diplatinum(II) complexes with complementary dipyridyl ligands in mixed solvents of dichloromethane and acetone at 60°C overnight [8]. These reactions typically proceed in nearly quantitative yields, demonstrating the efficiency and robustness of coordination-driven self-assembly. The preparation process of SCCs usually progresses from synthesizing simple small-molecule components to assembling metallacycles/metallacages, predominantly conducted in organic solvents [5]. For biomedical applications, additional steps are often required to enhance water solubility, either through chemical modification with hydrophilic groups or encapsulation within nanocarriers featuring hydrophilic surfaces [5].
Comprehensive characterization of ICSs requires multiple analytical techniques to confirm structure, purity, and composition:
Nuclear Magnetic Resonance (NMR) Spectroscopy: Multidimensional NMR techniques (¹H, ³¹P, DOSY) provide crucial information about molecular structure, symmetry, and purity [9] [8]. Coordination-induced chemical shifts in ¹H NMR (Îδ = 0.02-0.37 ppm for pyridyl protons) and ³¹P NMR (upfield shifts of approximately 4-5 ppm) serve as indicators of successful metal-ligand coordination [8]. DOSY NMR confirms that all proton signals possess the same diffusion coefficient, verifying the formation of discrete assemblies rather than mixtures [8].
Mass Spectrometry: Electrospray Ionization Time-of-Flight Mass Spectrometry (ESI-TOF-MS) enables the determination of molecular mass and identification of charge states through characteristic peaks corresponding to [M-nPFââ»]â¿âº species [9] [8]. This technique is particularly valuable for confirming the composition of large supramolecular architectures.
X-ray Crystallography: Single-crystal X-ray diffraction (SCXRD) provides unambiguous structural information at atomic resolution, allowing precise determination of cavity sizes, molecular geometries, and conformational details [9] [7]. SCXRD has been instrumental in characterizing structures ranging from simple metallacycles to complex metallacages and helicates.
Diagram 1: Experimental Workflow for ICS Research
ICSs demonstrate significant potential in biomedical applications, particularly in cancer therapy, diagnostic imaging, and drug delivery systems.
Supramolecular metal-based structures have emerged as novel anticancer agents with mechanisms of action distinct from classical small molecules [1] [6]. These compounds often exhibit cytotoxicity against human cancer cells comparable to or exceeding that of established chemotherapeutics like cisplatin [4]. For example, [PdâLâ]â´âº metallacages featuring bis-quinoline and bis-isoquinoline-based ligands demonstrated enhanced cytotoxicity against cancer cell lines compared to related pyridyl-based systems [4]. The antiproliferative effects of a self-assembled hexagonal Pd(II) macrocycle against a panel of cancer cell lines were found to be comparable to cisplatin [4]. The robustness and modular composition of supramolecular metal-based structures allow for the incorporation of different functionalities to enable active tumor targeting and stimuli-responsiveness [1].
The host-guest chemistry of 3D metallacages can be exploited to design novel drug delivery systems for anticancer chemotherapeutics [1] [6]. Metallacages feature internal cavities accessible to guest encapsulation, making them ideal for drug delivery applications [1]. Hydrophobic Pt(II)-organic cages can be formulated using nanoprecipitation techniques to enhance their bioavailability [4]. Encapsulation strategies can also reduce the toxicity of therapeutic agents; for instance, Q[10] encapsulation enhanced the pharmacokinetics of Ru(II) complexes, potentially enabling their use as antimicrobial agents [4]. The inherent flexibility of MOCs allows them to adapt their internal cavity to better accommodate guest molecules, exhibiting behavior analogous to the induced-fit mechanism of enzymes [7].
ICSs constitute ideal scaffolds for developing multimodal theranostic agents that combine therapeutic and diagnostic capabilities [1] [6]. The modular composition of supramolecular metal-based structures allows for the incorporation of imaging functionalities alongside therapeutic components [1]. Photoactive cages and capsules in which either the metal ion complexation or the bridging ligand are endowed with luminescence properties can be exploited for designing novel imaging agents, as well as for sensing and photoactivation in biological systems [1] [6]. Systems incorporating aggregation-induced emission (AIE) units like tetraphenylethylene (TPE) can overcome aggregation-caused quenching (ACQ) and the heavy atom effect on fluorescence imaging, facilitating the construction of efficient theranostic platforms [5].
Despite the significant progress in ICS research, several challenges remain to be addressed before widespread clinical translation can be realized.
The inherent hydrophobicity of many ICSs poses challenges for biomedical applications in aqueous physiological environments [5]. Most SCCs are constructed from hydrophobic components in organic solvents, leading to rapid aggregation and precipitation in biological systems [5]. Two primary strategies have been employed to enhance water solubility: (1) encapsulating hydrophobic ICSs within nanocarriers featuring hydrophilic surfaces, and (2) chemically modifying ICSs with hydrophilic groups or long chains to confer water solubility [5]. The surface charge, size, and chemical composition of these nanocarriers significantly influence their therapeutic efficacy, with optimal sizes typically ranging from 10-200 nm to leverage the enhanced permeability and retention (EPR) effect in tumor tissues [5].
The self-assembly of ICSs often involves heavy metal ions, raising concerns about potential toxicity and side effects during treatment [5]. Enhancing circulation stability and minimizing adverse effects are crucial for clinical translation [5]. Strategies to improve biosafety include developing stimuli-responsive systems that release therapeutic payloads specifically at target sites, functionalizing ICS surfaces with targeting moieties to enhance cellular selectivity, and exploring less toxic metal ions in coordination frameworks [5]. Research on silver and gold N-heterocyclic carbene (NHC) pillarplexes has shown promising antimicrobial and anticancer properties with differential bioactivity between the metal variants [4].
The construction of increasingly complex ICSs with precise functionality remains a synthetic challenge. Future research directions include the development of more sophisticated functionalization strategies, incorporation of stimuli-responsive elements for controlled guest release, and creation of hierarchical structures with multiple compartments for complex chemical transformations [9] [7]. The integration of ICSs with other materials, such as polymers, nanoparticles, and surfaces, will expand their application potential in targeted drug delivery, sensing, and catalysis [5].
Table 3: Research Reagent Solutions for ICS Construction
| Reagent Category | Specific Examples | Function in ICS Construction |
|---|---|---|
| Metal Acceptors | Pd(II), Pt(II), Cu(I), Fe(II), Ag(I) complexes | Provide structural nodes with specific coordination geometries |
| Organic Donors | Pyridyl-based ligands, catecholates, N-heterocyclic carbenes | Serve as linkers with defined angles and directions |
| Functional Ligands | Diarylethene, BODIPY, tetraphenylethylene (TPE) | Impart photoresponsive, fluorescent, or AIE properties |
| Solvent Systems | Dichloromethane, acetone, THF/water mixtures | Mediate self-assembly process and influence morphology |
| Counterions | PFââ», OTfâ» | Provide charge balance and influence solubility |
Diagram 2: Challenges and Strategies in ICS Development
Inorganic Cluster Supramolecular Compounds represent a rapidly advancing field with significant potential in biomedical applications. The precise control over structure and functionality afforded by coordination-driven self-assembly enables the design of sophisticated materials with tailored properties for drug delivery, cancer therapy, diagnostic imaging, and theranostics. While challenges related to water solubility, biocompatibility, and toxicity remain active areas of investigation, the unique characteristics of ICSsâincluding their structural precision, host-guest capabilities, and dynamic behaviorâposition them as promising candidates for next-generation pharmaceutical development. As research continues to address current limitations and explore new structural paradigms, ICSs are poised to make substantial contributions to advanced therapeutics and personalized medicine.
The field of supramolecular chemistry concerns chemical systems composed of discrete molecular components organized via non-covalent interactions, with metal-ligand coordination representing a particularly powerful subset due to its directionality and predictability [10]. Coordination-driven self-assembly has emerged as a foundational methodology for constructing sophisticated architectures from molecular building blocks, enabling the creation of both discrete supramolecular coordination complexes (SCCs) and extended metal-organic frameworks (MOFs) [11]. These metal-organic materials (MOMs) share the common design principle of metal nodes connected by organic ligands, yet diverge in their final structures and applications [11]. Understanding the fundamental bonding interactions and assembly pathways is crucial for advancing the rational design of inorganic cluster supramolecular compounds with tailored properties for applications ranging from catalysis to drug delivery [12].
This technical guide examines the core principles governing metal-ligand coordination and self-assembly pathways within the context of inorganic cluster supramolecular compound synthesis. We explore the historical foundations, quantitative bonding parameters, experimental methodologies, and characterization techniques essential for researchers engaged in the design and implementation of these complex systems.
The theoretical foundation for modern coordination chemistry was established by Alfred Werner in 1893 with his description of octahedral transition metal complexes, which explained how metal ions form bonds beyond what was necessary for charge neutrality [11]. This work originated the understanding that metal ions possess preferred coordination geometries that enable rational synthetic methodologies for installing specific ligands [11].
Supramolecular coordination chemistry gained significant momentum with pioneering work on molecular recognition in the 1960s by Pedersen, who discovered crown ethers capable of selectively chelating metal ions [11] [10]. This was followed by Lehn's development of cryptands and Cram's work on selective host-guest complexes, for which they shared the 1987 Nobel Prize in Chemistry [10]. These early investigations established that non-covalent interactions â including hydrogen bonding, metal coordination, hydrophobic forces, van der Waals forces, Ï-Ï interactions, and electrostatic effects â govern the spatial organization of molecular components into larger architectures [10].
The directional nature of metal-ligand coordination bonds provides a significant advantage over other non-covalent interactions for constructing well-defined architectures. The kinetic lability of these bonds enables self-correction during assembly, while the geometric preferences of metal ions (e.g., linear, square planar, tetrahedral, octahedral) allow for predictable structural outcomes when combined with appropriately designed organic ligands [11] [13].
Table 1: Common Metal Coordination Geometries in Supramolecular Assembly
| Metal Coordination Geometry | Typical Metal Ions | Ligand Arrangement | Resulting Architecture |
|---|---|---|---|
| Linear | Pt(II), Pd(II), Au(I) | 180° between ligands | Linear chains, macrocycles |
| Square Planar | Pt(II), Pd(II), Ni(II) | 90° between ligands | Squares, grids, 2D layers |
| Tetrahedral | Cu(I), Zn(II), Cd(II) | 109.5° between ligands | Tetrahedra, diamondoid networks |
| Octahedral | Fe(II), Ru(II), Co(III) | 90° between ligands | Trigonal prisms, cubes, 3D networks |
The selection of metal ion and organic ligand directly determines the structural outcome of the self-assembly process. Early work by Fujita and Stang in the 1990s demonstrated the rational design of supramolecular squares using Pd and Pt ions with linear bridging ligands [11]. This foundational approach has since been extended to create increasingly complex architectures, including polygons, polyhedra, prisms, and extended frameworks [11].
Self-assembly processes in supramolecular systems can proceed under thermodynamic or kinetic control, significantly impacting the reaction pathway and final product [13]. Thermodynamically controlled assemblies utilize reversible metal-ligand bonds that allow the system to explore multiple configurations before settling into the global energy minimum, which typically represents the most stable structure [13]. In contrast, kinetically controlled assemblies may form metastable intermediates that persist due to high activation energies for rearrangement.
The dynamic nature of metal-ligand coordination is essential for self-correction during assembly. Labile metal-ligand interactions enable initially formed, random oligomeric structures to rearrange into the thermodynamically favored product [13]. This spontaneous error-checking mechanism represents a significant advantage for constructing complex architectures with high fidelity.
Studies of helicate formation demonstrate that self-assembly often proceeds through a stepwise mechanism involving several kinetic intermediates [13]. These intermediates may possess altered coordination geometries or include coordinated solvent molecules that are eventually displaced during the structural self-correction process. In some systems, the addition of external agents such as guest molecules can trigger supramolecular transformations from one structure to another over extended time periods, providing insight into assembly mechanisms [13].
Cooperativity represents a fundamental principle in supramolecular assembly where the integration of constituent components creates emergent properties not present in the individual building blocks [13]. This phenomenon significantly impacts the stability and dynamic behavior of the resulting architectures.
Research on tris-catecholato complexes demonstrates this principle clearly. While mononuclear tris-catecholato complexes racemize rapidly via a trigonal twist mechanism, linking two such complexes in a helicate slows racemization by a factor of 100 due to mechanical coupling between metal centers [13]. This cooperative effect becomes even more pronounced in tetrahedral MâLâ structures, where racemization is not observed despite evidence of dynamic ligand and guest exchange [13]. Similar studies on Co²⺠benzimidazole complexes found the racemization rate of dinuclear helicates to be six orders of magnitude slower than their mononuclear analogs [13].
Table 2: Quantitative Studies of Cooperativity in Supramolecular Assemblies
| Assembly Type | Structural Feature | Dynamic Process | Rate Change vs. Monomer |
|---|---|---|---|
| MâLâ helicate | Two linked tris-catecholate complexes | Racemization | 100Ã slower |
| MâLâ tetrahedron | Four linked tris-catecholate complexes | Racemization | Not observed |
| CoâLâ helicate | Dinuclear benzimidazole complex | Racemization (dissociative mechanism) | 10â¶Ã slower |
The construction of supramolecular architectures requires careful design of molecular building blocks with specific geometric and electronic properties. Linear metal-organic ligands with specific sequences can be designed to facilitate the self-assembly of metallo-supramolecules with increasing complexity [14]. These building blocks typically consist of organic ligands with multiple binding sites oriented with specific angularity and metal ions with well-defined coordination preferences [11].
Recent advances have demonstrated the synthesis of linear building blocks through coordination between terpyridine ligands and Ru(II), creating metal-organic ligands with high stability [14]. Subsequent self-assembly with weaker coordinating but highly reversible metal ions such as Zn(II), Fe(II), and Cd(II) enables the formation of discrete 2D fractal architectures ranging from simpler structures (C1, 3,360 Da) to highly complex systems (C5, 38,066 Da) with precisely controlled shapes and sizes [14].
Three primary synthetic approaches for preparing terpyridine-Ru(II) metal-organic ligands have been developed:
The Sonogashira approach particularly enables the synthesis of longer asymmetric metal-organic building blocks with rigid linkages, reminiscent of protection and deprotection strategies in peptide synthesis [14].
Rigorous characterization of supramolecular assemblies presents significant challenges due to their large size, complex connectivity, and dynamic behavior [13]. Multiple complementary techniques are typically required to unambiguously determine structure and assembly properties.
Mass spectrometry techniques, particularly electrospray ionization mass spectrometry (ESI-MS) and traveling wave ion mobility-mass spectrometry (TWIM-MS), provide essential information about molecular weight, composition, and isomeric purity [14]. ESI-MS can reveal continuous charge states with well-resolved isotope patterns consistent with theoretical distributions, while TWIM-MS with narrow drift time distribution confirms the absence of isomeric structures [14].
Multi-dimensional NMR spectroscopy, including ¹H, COSY, and NOESY experiments, offers insights into molecular symmetry and structural features [14]. For instance, in zinc-coordinated terpyridine assemblies, characteristic downfield shifts of 3',5' protons and upfield shifts of 6,6" protons confirm successful metal coordination [14]. Similarly, ¹â¹F NMR can reveal symmetry information through discrete chemical environments [14].
Table 3: Characterization Techniques for Supramolecular Assemblies
| Technique | Information Obtained | Key Observations |
|---|---|---|
| ESI-MS | Molecular weight, composition | Continuous charge states, isotope patterns |
| TWIM-MS | Structural isomers, shape | Drift time distribution |
| ¹H NMR | Molecular symmetry, ligand environment | Characteristic proton shifts |
| ¹â¹F NMR | Symmetry, chemical environments | Discrete sets of peaks |
| 2D COSY/NOESY | Spatial relationships, connectivity | Through-bond and through-space correlations |
The following diagram illustrates the fundamental pathway for coordination-driven self-assembly of supramolecular architectures:
The comprehensive characterization of supramolecular assemblies requires a multi-technique approach as depicted below:
Table 4: Essential Research Reagents in Supramolecular Coordination Chemistry
| Reagent Category | Specific Examples | Function in Assembly | Key Characteristics |
|---|---|---|---|
| Metal Salts | Cd(NOâ)â, Zn(II) salts, Fe(II) salts, Pt(II) complexes, Pd(II) complexes | Geometric nodes with specific coordination preferences | Variable coordination geometry, lability, oxidation state |
| Nitrogen-based Ligands | 4,4'-bipyridine, terpyridine, phenanthroline | Linear or angular spacers between metal nodes | Rigidity, binding affinity, angularity |
| Carboxylate-based Ligands | Terephthalate, trimesic acid, 1,3,5-benzenetricarboxylic acid | Bridging ligands for extended frameworks | Multiple binding modes, charge balance |
| Macrocyclic Hosts | Cucurbiturils, cyclodextrins, calixarenes | Molecular recognition elements | Preorganized cavities, host-guest chemistry |
| Template Molecules | Chloride ions, organic cations, neutral aromatics | Structure-directing agents | Complementary size/shape to assembly cavity |
| Solvent Systems | CHClâ/MeOH mixtures, aqueous ethanol, DMF | Medium for self-assembly process | Polarity, coordinating ability, solubility |
The rational design of supramolecular architectures through metal-ligand coordination and self-assembly pathways represents a sophisticated synthetic methodology at the interface of inorganic and supramolecular chemistry. Fundamental understanding of coordination geometries, thermodynamic versus kinetic control, cooperativity effects, and characterization techniques enables researchers to construct complex functional systems with precision. The continued development of this field promises advanced materials for biomedical applications, catalysis, sensing, and molecular electronics, driven by increasingly sophisticated design principles and experimental methodologies. As the field progresses, integration of biological inspiration with synthetic expertise will likely yield ever more complex and functional supramolecular systems.
The rational design of inorganic cluster supramolecular compounds hinges on the precise selection and combination of two fundamental components: transition metal nodes and multidentate organic linkers. These building blocks dictate the structural topology, stability, and functional properties of the resulting frameworks, which find applications in catalysis, drug delivery, bioimaging, and theranostics. This technical guide provides an in-depth analysis of these core components, detailing their classifications, properties, and the experimental methodologies that enable the synthesis of advanced supramolecular architectures. Framed within current research on inorganic cluster supramolecular chemistry, this whitepaper serves as a foundational resource for researchers and drug development professionals engaged in the design of functional coordination complexes.
Inorganic cluster supramolecular compounds are a class of materials constructed from metal-based nodes connected by organic linkers via coordination bonds. The field represents a convergence of coordination chemistry and supramolecular science, focusing on the design of discrete or extended architectures with defined geometries and properties. [6] The robustness and modular nature of these structures allow for the incorporation of diverse functionalities, making them ideal scaffolds for developing multimodal theranostic agents and novel drug delivery systems. [6] The design principles are largely governed by the "node-and-linker" approach, where the coordination geometry of the metal cluster (node) and the binding topology of the organic ligand (linker) determine the overall structure and, consequently, the material's functional characteristics. [15] [16]
Transition metal clusters, particularly octahedral clusters of molybdenum or rhenium, offer unique photophysical and redox properties, while multidentate organic linkers provide the structural directionality and chemical functionality. [17] [18] The synergy between these components enables the creation of materials with tailored properties for specific applications, from photodynamic therapy to catalysis. [17] [19] This guide systematically outlines the key building blocks, their classifications, and the experimental protocols for their assembly, providing a comprehensive resource for synthesis research in this rapidly evolving field.
Transition metal nodes, often referred to as Secondary Building Units (SBUs), are inorganic clusters that serve as the structural vertices of supramolecular frameworks. These nodes can range from single metal ions to polynuclear metal clusters, with their specific coordination geometry dictating the overall topology of the resulting network. [15]
Table 1: Key Classes of Transition Metal Cluster Nodes and Their Characteristics
| Cluster Type | Composition | Key Structural Features | Photophysical Properties | Representative Applications |
|---|---|---|---|---|
| Octahedral Halide Clusters | [M6L14]n- (M = Mo, W; L = halogen) [17] | Nanometer-sized metallic Mo(II) aggregates; 8 inner ligands, 6 labile apical ligands [17] | Red/NIR phosphorescence; high singlet oxygen quantum yields (e.g., [Mo6I8(CF3COO)6]2- ISC: 1.7 ps) [17] | Photodynamic/Radiodynamic Therapy [17] |
| Octahedral Cyano Clusters | [M6L14]n- (M = Mo, Re; L = chalcogen) [18] | [M6L14]n- units, 1< n <8 [18] | Highly emissive in red-NIR (PLQY up to 0.23) [18] | Luminescent dyes in Liquid Crystals [18] |
| Zirconium-based SBUs | e.g., Zr6O4(OH)4 (in UiO-66) [19] | 6 Zr atoms in an octahedron; high connectivity | High chemical and thermal stability | Catalysis, Drug Delivery [19] |
| Paddle-wheel SBUs | e.g., Cu2(RCOO)4 | Dinuclear cluster with four bridging linkers | Open metal sites after activation | Gas storage, Separation |
A prominent class of nodes is the octahedral molybdenum cluster. For instance, the [Mo6I8]n+ core consists of six molybdenum atoms forming an octahedron, stabilized by eight strongly bonded inner iodide ligands and six labile apical ligands that can be functionalized. [17] These clusters display attractive photophysical properties, including excitation by UV-blue light, ultrafast intersystem crossing (e.g., 1.7 ps for [Mo6I8(CF3COO)6]2-), and long-lived triplet states (hundreds of µs) in solids. [17] Their phosphorescence in the red/near-IR region, with large Stokes shifts and high quantum yields, makes them excellent photosensitizers for producing singlet oxygen (O2(1Îg)), a key mediator in photodynamic therapy (PDT) and antimicrobial photoinactivation. [17]
The electronic properties of these metal clusters, such as the energy of their triplet states (approximately 1.9 eV for Mo6 clusters), must be higher than the energy required to excite molecular oxygen to its singlet state (0.98 eV) for efficient energy transfer and singlet oxygen production. [17] Their inorganic nature also makes them less prone to photobleaching compared to organic dyes like porphyrins, and they do not experience self-quenching of excited states at high concentrations, enhancing their utility in solid-state applications or as concentrated agents. [17]
Organic linkers are multidentate bridging ligands that connect metal nodes, defining the framework's geometry, pore size, and chemical environment. The flexibility, length, and functional groups of the linker directly influence the structural and chemical properties of the final supramolecular compound. [20]
Table 2: Classification of Multidentate Organic Linkers for Supramolecular Compounds
| Linker Class | Functional Group | Coordination Mode & Characteristics | Impact on Framework Properties | Example Applications |
|---|---|---|---|---|
| Carboxylate Linkers | -COOâ» | Chelates/bridges metal ions; strong coordination; H-bond donor/acceptor [20] | High surface area, uniform pores, high crystallinity [20] | Gas adsorption, Catalysis, Metal removal [20] |
| Nitrogen-based Linkers | Heterocyclic N (e.g., pyridyl, imidazole) | Directional coordination; tunable electronic properties [20] | Tunable pore chemistry, structural diversity [20] | Molecular sieving, Catalysis, Pollutant capture [20] |
| Phosphorous-based Linkers | -PO3H / -PO3â» | Forms strong P-O-M bonds [20] | High chemical/thermal stability; polar porosity [20] | Proton conduction, Luminescent sensing [20] |
| Sulfonic Acid Linkers | -SO3H | Strong Brønsted acid site; polar, electron-rich [20] | Hydrophilicity; acidic catalysis in pores [20] | Acid-catalyzed reactions [20] |
| Flexible Multidentate Linkers | Mixed N/O donors (e.g., pyrazole-carboxylate) | Adaptable conformation; variable binding modes [21] | Structural dynamism; formation of low-dimension structures (e.g., 1D chains) [21] | Luminescent materials [21] |
Carboxylate-based linkers are among the most extensively used due to their strong coordination to metal ions and ability to form rigid, highly porous frameworks. [20] For example, 1,4-benzenedicarboxylic acid (terephthalic acid, H2bdc) is a fundamental linker in iconic MOFs like MOF-5. [15] The rigidity of the aromatic backbone in such linkers is key to achieving permanent porosity.
Nitrogen-based linkers, particularly heterocyclic types like pyridyl, imidazole, and their derivatives, offer different coordination geometries and electronic characteristics compared to carboxylates. [20] They are classified into heterocyclic azine N-based linkers (e.g., pyrazine, bipyridyl), heterocyclic-azole N-based linkers (e.g., imidazole, triazole), noncyclic N-based linkers, and ionic N-based linkers. [20] The choice of nitrogen linker significantly influences the framework's stability, pore environment, and application suitability. [20]
The functionalization of linker backbones allows for the fine-tuning of host-guest chemistry within the framework's pores. For instance, sulfonic acid groups can introduce strong Brønsted acidity for catalysis, while hydroxy groups can enhance CO2 capture and facilitate proton conduction. [20] Fluorinated linkers impart hydrophobicity and stability, and their highly polar C-F bonds can improve interactions with gases like CO2. [20]
The synthesis of supramolecular compounds based on transition metal clusters and organic linkers requires careful control of reaction conditions to achieve the desired crystalline products.
Solvothermal Synthesis: This is the most common method for growing high-quality single crystals suitable for structure determination by X-ray crystallography. [15] It typically involves dissolving the metal salt and organic linker in a suitable solvent (e.g., N,N-diethylformamide, water, acetonitrile) in a sealed vessel and heating the mixture to temperatures above the solvent's boiling point for several hours to days. [15] The elevated temperature and pressure enhance the solubility and reactivity of the precursors, facilitating slow and controlled crystal growth. For example, the synthesis of complex [Cu(L1)(2,2âbipy)]2n·3nH2O (1) was achieved by reacting the flexible ligand 1,1´-methylenebis(5-methyl-pyrazole-4-carboxylic acid) (H2L1) and coligand 2,2'-bipyridyl with Cu(OAc)2·H2O under specific, though undisclosed, reaction conditions. [21]
Microwave-Assisted Synthesis: This technique uses microwave irradiation to rapidly nucleate MOF crystals, reducing synthesis times from days to minutes or seconds. [15] It often yields smaller, more uniform crystals (micron-scale) compared to conventional solvothermal methods and is useful for high-throughput screening. [15]
Mechanochemical Synthesis: A solvent-free approach that involves grinding metal precursors and organic linkers using a ball mill. [15] This method is scalable and environmentally friendly, producing materials like Cu3(BTC)2 in quantitative yield. The morphology of the solvent-free synthesized product can be identical to the industrially produced material (e.g., Basolite C300). [15] The addition of small amounts of ethanol can reduce structural defects. [15]
Electrophoretic Deposition (EPD) for Film Formation: For applications requiring thin films, such as coatings for medical devices, EPD can be employed. Kirakci et al. demonstrated the electrophoretic deposition of layers of octahedral molybdenum cluster complexes onto surfaces, creating coatings effective for mitigating pathogenic bacterial biofilms under blue light. [17]
To enhance the stability, dispersibility, and biodistribution of cluster complexes for biological applications, post-synthetic modifications are often employed:
Table 3: Key Research Reagent Solutions for Supramolecular Compound Synthesis
| Reagent/Material | Function | Specific Examples |
|---|---|---|
| Metal Salts | Source of metal ions/clusters | Cu(OAc)2·H2O [21], ZrCl4, MoCl2 |
| Carboxylate Linkers | Primary bridging ligands for framework construction | 1,4-benzenedicarboxylic acid (H2BDC) [15], 1,3,5-benzenetricarboxylic acid (H3BTC) |
| Nitrogen-based Linkers | Provide directional coordination and tunable functionality | 2,2'-bipyridyl (2,2-bipy) [21], 4,4'-bipyridine, imidazole derivatives |
| Solvents | Medium for solvothermal synthesis; influences crystal growth | N,N-Diethylformamide (DEF), Water, Acetonitrile [15] |
| Modulators | Competitive coordinating agents to control crystal growth | Acetic acid, Trifluoroacetic acid |
| Functionalization Reagents | Post-synthetic modification of clusters or linkers | PEG derivatives [17], Chitosan [17], Targeting peptides |
| Nanocarriers | Improve biocompatibility and biodistribution for biological apps | PLGA [17], HPMA copolymer [17] |
| Stannane, butyltriiodo- | Stannane, butyltriiodo-, CAS:21941-99-1, MF:C4H9I3Sn, MW:556.54 g/mol | Chemical Reagent |
| Urea, (p-hydroxyphenethyl)- | Urea, (p-hydroxyphenethyl)- | Urea, (p-hydroxyphenethyl)- is a chemical for research (RUO). It is not for human, veterinary, or household use. Explore its value as a urease inhibitor and antimicrobial agent. |
The following diagram illustrates the logical relationship between the choice of building blocks, the synthesis process, the resulting structural properties, and the eventual applications of the supramolecular compounds.
In the field of inorganic cluster supramolecular compound synthesis, architectural control is a fundamental objective, determining the physical properties and ultimate application of the materials formed. This control is exercised primarily through two interdependent elements: the coordination geometry of the metal nodes and the molecular design of the organic linkers. The metal nodes, with their specific coordination preferences dictated by electronic structure and the nature of the coordinating atoms, act as the vertices of the final structure. The organic linkers, through their length, functionality, and directional bonding character, define the edges and dictate the spatial arrangement of these vertices. This whitepaper provides an in-depth technical guide to the principles and methodologies for achieving precise architectural control in the synthesis of supramolecular assemblies, with a specific focus on systems involving inorganic clusters.
The geometry around a metal center is a primary determinant of the final network topology in supramolecular assemblies. The coordination geometry is influenced by the metal ion's character, oxidation state, and the chemical nature of the ligands provided by the organic linkers.
Table 1: Common Metal Center Coordination Geometries and Their Structural Influence
| Metal Ion Example | Coordination Number | Geometry | Ideal Bond Angle | Representative Cluster Topology |
|---|---|---|---|---|
| Cu(I), Ag(I) [22] | 2 | Linear | 180° | Chains, 1D networks |
| Cu(II), Pt(II) [22] | 4 | Square Planar | 90° | 2D grid layers |
| Zn(II), Cd(II) | 4 | Tetrahedral | 109.5° | Diamondoid, 3D networks |
| Cu(II) | 4 + 2 | Jahn-Teller Distorted Octahedral | Axial: ~180°; Equatorial: ~90° | Paddlewheel units in MOFs |
| Zr(IV) | 6, 8, 12 | Octahedral, Square Antiprismatic | Varies | High-connectivity, stable frameworks |
The use of organometallic complexes as primary linkers, rather than simple organic molecules, represents an advanced strategy for architectural control. For instance, metalated quinonoid complexes, such as [Cp*M(ηâ´-benzoquinone)] (where Cp* = Câ
Meâ
, M = Rh, Ir), can act as supramolecular building blocks [22]. These units possess their own intrinsic coordination preferences from the Cp*M moiety and can further coordinate through their oxygen or sulfur atoms to secondary metal ions (e.g., Cu(I), Ag(I), Pt(II)) [22]. This creates a hierarchical assembly process where the coordination geometry at both the organometallic core and the secondary metal node directs the formation of complex supramolecular architectures.
The organic linker is the architectural blueprint in supramolecular assembly. Its design directly controls the network's metrics, topology, and chemical functionality.
Achieving architectural control requires precise and reproducible synthetic and analytical methodologies.
The following protocol, adapted from procedures for synthesizing networks with organometallic linkers and supramolecular colloids, outlines a general solvothermal approach [22] [25].
Objective: To synthesize a supramolecular coordination network using an organometallic quinonoid linker and a secondary metal ion.
Materials:
[Cp*Ir(ηâ´-benzoquinone)])Procedure:
The following workflow diagram illustrates the key stages of this synthesis and the subsequent characterization process.
Rigorous characterization is essential to confirm that the synthesized structure matches the designed architecture.
Table 2: Essential Reagents and Materials for Supramolecular Compound Synthesis
| Reagent/Material | Function & Role in Architectural Control |
|---|---|
| Metal Salts (e.g., Cu(CFâSOâ)â, AgNOâ, Zn(NOâ)â) | Source of metal ions that act as structural nodes; the choice of metal and counter-ion influences coordination geometry and network solubility [22]. |
Organometallic Linkers (e.g., [Cp*M(ηâ´-benzoquinone)] M=Rh, Ir) |
Function as pre-designed, stable spacers with defined coordination vectors, generating novel architectures not accessible with purely organic linkers [22]. |
| Functionalized Organic Linkers (e.g., Schiff bases, BTA derivatives) | Provide structural metrics (length, angle) and introduce specific intermolecular interaction sites (H-bonding) to guide and stabilize assembly [25] [24]. |
| Anhydrous, Degassed Solvents (e.g., CHâClâ, MeCN, DMF) | Reaction medium; purity is critical for preventing hydrolysis or oxidation of sensitive metal precursors and linkers, ensuring reproducible crystal growth [22] [25]. |
| Silica-based Colloids | Functionalizable substrates for creating supramolecular colloids; their surface can be grafted with moieties like BTA to study and control particle-level assembly [25]. |
| Photo-labile Protecting Groups (e.g., o-nitrobenzyl (NVOC)) | Used to block specific interactions (e.g., H-bonding) on linkers or colloids. Controlled deprotection with UV light triggers on-demand assembly, providing temporal control [25]. |
| 3-Methyl-5-phenylbiuret | 3-Methyl-5-phenylbiuret|High-Purity Research Chemical |
| 1,2,4-Triazine, 5-phenyl- | 1,2,4-Triazine, 5-phenyl-, CAS:18162-28-2, MF:C9H7N3, MW:157.17 g/mol |
Precise architectural control in inorganic cluster supramolecular compounds is not a singular achievement but the result of a holistic design strategy. It requires the synergistic integration of metal node engineeringâexploiting specific coordination geometriesâand sophisticated linker designâcontrolling metrics, functionality, and supramolecular recognition. The experimental protocols, centered on controlled synthesis and multi-technique characterization, are essential for verifying architectural outcomes and understanding assembly mechanisms. As the field progresses, the incorporation of stimuli-responsive and mechanically interlocked components will further enhance our ability to create dynamic, functional supramolecular materials with bespoke architectures for advanced applications in catalysis, drug delivery, and sensing.
Metallacages represent a prominent class of three-dimensional supramolecular architectures formed via coordination-driven self-assembly of metal acceptors and organic ligands. [26] [27] These structures have emerged as powerful hosts in supramolecular chemistry due to their well-defined cavities, tunable sizes and shapes, and unique electronic properties. The integration of metal centers and organic components creates hybrid materials that combine advantages of both inorganic and organic systems, including structural robustness, rich photophysical properties, and synthetic versatility. [28] [29] The development of metallacage-based host-guest systems represents a growing frontier in inorganic cluster supramolecular chemistry, with particular relevance to applications in molecular recognition, drug delivery, and catalysis. [29] [27]
Host-guest chemistry within metallacage cavities is governed by multiple complementary interactions, including Ï-Ï stacking, hydrophobic effects, hydrogen bonding, and van der Waals forces. [28] [30] The confined nanospaces within these structures mimic enzyme binding pockets, enabling selective encapsulation of guest molecules through complementary size, shape, and functionality. [29] This review comprehensively examines the fundamental principles, quantitative binding relationships, experimental methodologies, and emerging applications of host-guest systems based on metallacage cavities, with particular emphasis on recent advances in the field.
The construction of metallacages relies on directional bonding approaches where metal centers with specific coordination geometries connect organic ligands to form discrete, three-dimensional structures. [29] The most common design strategy utilizes cis-protected square planar metal complexes (particularly Pt(II) and Pd(II)) which provide 90° turns that promote cage closure. [29] [27] These metal centers typically coordinate to pyridyl donors on the organic ligands, forming stable metal-ligand bonds that define the cage architecture.
The geometry of the resulting metallacage is determined by the combination of metal coordination geometry and ligand topology. For instance, tetratopic porphyrin-based ligands combined with Pt(II) corners can form box-like structures, while tritopic ligands may generate triangular or tetrahedral architectures. [28] [29] The symmetry interaction model provides a complementary framework for predicting assembly outcomes, particularly for systems involving higher-symmetry components. [29] This approach considers the point group symmetries of both the metal corners and organic ligands to determine the most stable assembly product.
A key advancement in metallacage design has been the development of multicomponent self-assembly strategies, where multiple different organic ligands are incorporated into a single structure. [28] This approach enables finer control over cavity size, shape, and functionality, allowing for customization of host-guest properties. For example, porphyrin-based metallacages can be constructed using tetrapyridyl porphyrin faces connected by tetracarboxylic pillars via platinum(II) coordination bonds, resulting in box-shaped structures with ideal dimensions for encapsulating planar aromatic guests. [28]
The molecular recognition properties of metallacage cavities arise from their confined nanospaces and specific interior functionalities. The host-guest binding can be described by the equilibrium H + G â HG, where the association constant (K) quantifies the binding strength. [30] These interactions are typically fast-exchange processes on the NMR timescale, allowing for detailed thermodynamic and kinetic analysis. [26] [28]
Multiple factors contribute to guest encapsulation in metallacages:
The confined environment of metallacage cavities can significantly alter guest properties, including enhanced stability, modified reactivity, and changed spectroscopic signatures. [30] This confinement effect mimics enzymatic binding pockets and enables applications in stabilization of reactive species and control of chemical reactions.
Table 1: Experimentally Determined Association Constants for Metallacage-Guest Complexes
| Metallacage System | Guest Molecule | Association Constant (Kâ) | Experimental Conditions | Primary Interactions |
|---|---|---|---|---|
| PDI-based Metallacage 4 [26] | Pyrene (G1) | (2.63 ± 0.05) à 10³ Mâ»Â¹ | CDâCN, NMR Titration | Ï-Ï Stacking, Hydrophobic |
| PDI-based Metallacage 4 [26] | Coronene (G2) | (1.41 ± 0.07) à 10â´ Mâ»Â¹ | CDâCN, NMR Titration | Ï-Ï Stacking, Hydrophobic |
| Porphyrin-based Metallacage 4 [28] | Coronene (G6) | 2.37 à 10â· Mâ»Â¹ | CHâCN/CHClâ (9:1) | Ï-Ï Stacking, Charge Transfer |
| PDI-based Metallacage 4 [26] | Alizarin Red S (G3) | Significant upfield shifts in ¹H NMR | CDâCN | Host-Guest Complexation |
| PDI-based Metallacage 4 [26] | Methyl Orange (G4) | Significant upfield shifts in ¹H NMR | CDâCN | Host-Guest Complexation |
Table 2: Structural Parameters of Representative Metallacages and Their Cavities
| Metallacage | Dimensions | Cavity Volume | Metal-Ligand Bonds | Structural Features |
|---|---|---|---|---|
| PDI-based Metallacage 4 [26] | 1.65 à 1.41 à 0.98 nm³ | ~1.1 nm³ | Eight Pt(II)-coordination bonds | Two PDI faces, Two tetracarboxylic pillars, Dihedral angle: 41.9° |
| Porphyrin-based Metallacage 4 [28] | 14.9 à 12.0 à 8.3 à ³ | ~1.48 nm³ | Eight Pt atoms with N-Pt-O angles of 82.3-83.5° | Two porphyrin panels, Parallel orientation, Distance between panels: 8.1 à |
| Porphyrin-based Box [28] | Not specified | Not specified | Pt(II)-pyridyl/ carboxylate | Two large windows for guest entry, Nanochannel formation in solid state |
The quantitative data presented in Tables 1 and 2 reveal several important trends in metallacage host-guest chemistry. First, binding affinity correlates strongly with guest size and planarity, with larger, more planar aromatic systems (e.g., coronene) exhibiting significantly higher association constants than smaller analogues (e.g., pyrene). [26] [28] Second, metallacages with porphyrin panels generally show higher binding affinities than those with PDI faces, likely due to enhanced Ï-Ï interactions with aromatic guests. Third, the structural parameters of the cage cavity directly determine the size range of guests that can be effectively encapsulated.
The preparation of metallacages typically involves self-assembly under mild conditions to preserve the metal-coordination bonds. Recent advances have demonstrated the efficacy of photo-induced copolymerization for creating metallacage-crosslinked networks while maintaining cage integrity. [26]
Protocol 1: General Procedure for Metallacage Synthesis via Self-Assembly [26] [28]
Protocol 2: Photo-induced Copolymerization for Metallacage-Crosslinked Networks [26]
Protocol 3: Determination of Association Constants by NMR Titration [26] [30]
Protocol 4: Host-Guest Complexation for Planar Dyes [26]
Table 3: Key Reagents and Materials for Metallacage Host-Guest Studies
| Reagent/Material | Function/Application | Key Characteristics | Representative Examples |
|---|---|---|---|
| cis-(PEtâ)âPt(OTf)â [26] [28] | Metal acceptor for coordination-driven self-assembly | Square planar geometry, protected cis sites, moisture-sensitive | Pt(II) corners for box-shaped metallacages |
| Tetrapyridyl PDI Ligands [26] | Electron-deficient face for metallacage construction | Strong visible absorption, fluorescence, planar structure | Component of photoresponsive metallacages |
| Tetrapyridyl Porphyrin Ligands [28] | Photosensitizing face for metallacage assembly | Strong Soret band, singlet oxygen generation, redox activity | Component of porphyrin-based metallacages |
| Tetracarboxylic Pillar Ligands [26] [28] | Structural pillars connecting cage faces | Variable lengths, functionalizable with polymerizable groups | Control cavity size and functionality |
| Polycyclic Aromatic Hydrocarbons [26] [28] | Model guests for binding studies | Planar, varying sizes, strong Ï-Ï interactions | Pyrene, coronene, perylene derivatives |
| Planar Dyes [26] | Charged guests for complexation studies | Ionic, visible absorption, industrial relevance | Alizarin red S, methyl orange, methylene blue |
| Deuterated Solvents [26] [28] | Medium for NMR characterization | Isotopically pure, minimal water content | CDâCN, CDClâ, DMSO-dâ |
| Diethenyl ethanedioate | Diethenyl Ethanedioate|C6H6O4|Research Chemical | Research-grade Diethenyl Ethanedioate (C6H6O4). This product is For Research Use Only (RUO) and is not intended for personal use. | Bench Chemicals |
| Cobalt--dysprosium (1/3) | Cobalt--dysprosium (1/3), CAS:12200-33-8, MF:CoDy3, MW:546.43 g/mol | Chemical Reagent | Bench Chemicals |
Metallacages functionalized through host-guest chemistry have enabled diverse applications, particularly in environmental remediation and stimuli-responsive systems. PDI-based metallacage-crosslinked networks demonstrate excellent capabilities for photocatalytic water decontamination, effectively eliminating organic pollutants and bacterial contaminants through singlet oxygen generation. [26] These materials can be reused multiple times without significant loss of activity, making them practical for environmental applications.
Porphyrin-based metallacages exhibit remarkable singlet oxygen generation capabilities, enabling the development of photooxidation-responsive host-guest systems. [28] These systems can oxidize encapsulated anthracene derivatives to their endoperoxides, triggering guest release. By selecting anthracene guests whose endoperoxides can revert upon heating, fully reversible encapsulation-release systems can be constructed for controlled delivery applications. [28]
The future development of metallacage-based host-guest systems will likely focus on increasing complexity and functionality, including low-symmetry cages with anisotropic cavities, environmentally responsive systems, and integration with other materials. [29] Computational approaches are increasingly valuable for predicting host-guest compatibility and guiding synthetic efforts. [29] As these systems mature, they hold significant promise for advanced applications in drug delivery, chemical sensing, and green chemistry.
Diagram 1: Metallacage assembly and host-guest chemistry workflow. The process begins with self-assembly of metal acceptors and organic ligands to form 3D metallacages with defined cavities, followed by encapsulation of guest molecules, leading to functional applications.
Diagram 2: Molecular recognition principles in metallacage host-guest chemistry. Multiple complementary factors contribute to selective guest encapsulation within metallacage cavities.
Precursor-based synthesis, particularly through calcination, represents a foundational methodology in the fabrication of advanced catalytic nanomaterials. This approach leverages molecular-level control to dictate the structural, morphological, and compositional properties of the final inorganic material, enabling precise tuning of catalytic performance. The process fundamentally involves the thermal transformation of molecular or supramolecular precursorsâoften inorganic complexes or organic-inorganic hybridsâinto target metal oxide nanomaterials under controlled atmospheres [32] [33]. Within the broader context of inorganic cluster supramolecular compound research, this methodology bridges the gap between molecular chemistry and materials science, allowing the transfer of structural information from the precursor to the functional nanomaterial [34].
The calcination pathway offers distinct advantages for catalytic applications, including control over phase composition, surface area, particle size, and porosityâall critical parameters governing catalytic activity, selectivity, and stability. This technical guide provides an in-depth examination of precursor-based synthesis routes, with a specific focus on calcination processes for developing advanced catalytic nanomaterials, drawing upon recent advances in supramolecular and nanomaterial chemistry.
The thermal decomposition of precursors during calcination follows complex solid-state reaction pathways influenced by multiple parameters. Understanding these fundamentals is essential for rational design of catalytic nanomaterials.
Precursor compounds serve as molecular reservoirs that define the stoichiometry and spatial distribution of elements in the final material. The selection of precursor classes includes:
The decomposition pathway of any precursor is governed by the characteristics of the organic moiety, the gaseous atmosphere (inert, reducing, or oxidizing), and the thermal profile employed during calcination [33].
The transformation from precursor to metal oxide involves complex thermodynamic and kinetic processes:
For zirconia nanomaterials, for instance, calcination temperature directly controls crystalline phase: pure monoclinic phase is stable up to 1100°C, tetragonal phase forms between 1100â2370°C, and cubic phase exists above 2370°C [32]. Stabilization of typically high-temperature phases at room temperature can be achieved through precursor design and controlled calcination.
Table 1: Classification of Precursor Types for Nanomaterial Synthesis
| Precursor Category | Specific Examples | Decomposition Products | Key Advantages |
|---|---|---|---|
| Inorganic Salts | ZrOClâ·8HâO, ZrClâ, Co(NOâ)â | ZrOâ, CoâOâ | Low cost, simple processing |
| Carboxylate Complexes | Cobalt oxalate, citrate, malonate | CoâOâ, CoO, metallic Co | Controlled stoichiometry, mixed metal capability |
| Alkoxides | Zr(Câ HâOâ)â | ZrOâ | High purity, low decomposition temperatures |
| Supramolecular Hybrids | {[L][HgIâ]}, {[L][CoClâ]â} | Metal oxides, chalcogenides | Molecular-level structure control |
Oxalate Route for Cobalt-Based Nanomaterials:
Citrate-Based Sol-Gel Method:
Organic-Inorganic Hybrid Template Synthesis:
Calcination transforms precursors to active nanomaterials through controlled thermal treatment:
Standard Calcination Protocol:
Atmosphere Control:
Table 2: Optimization of Calcination Parameters for Selected Nanomaterials
| Target Material | Precursor | Optimal Calcination Temperature | Atmosphere | Holding Time | Resulting Properties |
|---|---|---|---|---|---|
| CoâOâ nanoparticles | Cobalt oxalate dihydrate | 500°C | Static air | 2 hours | Spinel structure, 20-30 nm size [33] |
| ZrOâ nanoparticles | Zirconium oxychloride | 400°C | Air | 2 hours | Mixed phase (monoclinic/tetragonal), 8-10 nm size [32] |
| CoO nanoparticles | Cobalt oxalate | 500°C | Air | 4 hours | Rock salt structure, anisotropic morphology [33] |
| CoFeâOâ spinel | Oxalate precursor | 600°C | Air | 5 hours | Ferrimagnetic, 15-25 nm size [33] |
Comprehensive characterization establishes structure-property relationships essential for catalytic applications:
Calcined nanomaterials exhibit diverse catalytic applications governed by their structural and electronic properties:
Organic-inorganic hybrid supramolecules demonstrate exceptional photocatalytic performance. For instance, compound {[(L)(CuâIâ)]·[CuIâ]CHâCN}â (3) achieves 92.22% tetracycline degradation efficiency under optimal conditions (10 mg catalyst, pH = 7) [34]. The photocatalytic efficiency remains above 86% after four cycles, demonstrating excellent recyclability. The mechanism involves photo-generated electron-hole pairs that initiate radical formation, decomposing organic pollutants.
Zirconia-based nanomaterials function as efficient catalysts and catalyst supports for various reactions due to their tunable acid-base properties and thermal stability. The catalytic activity directly correlates with the crystalline phase and surface area, which are predetermined by precursor selection and calcination parameters [32].
Cobalt oxide nanomaterials derived from carboxylate precursors exhibit promising electrocatalytic activity for oxygen evolution reaction (OER) and oxygen reduction reaction (ORR), with performance dependent on calcination-induced morphology and surface defect chemistry [33].
Table 3: Key Reagent Solutions for Precursor-Based Nanomaterial Synthesis
| Reagent/Chemical | Function/Purpose | Application Example |
|---|---|---|
| Zirconium oxychloride octahydrate (ZrOClâ·8HâO) | Zirconia precursor | Hydrothermal synthesis of ZrOâ nanoparticles [32] |
| Cobalt acetate tetrahydrate (Co(CHâCOO)â·4HâO) | Cobalt source for carboxylate precursors | Synthesis of cobalt oxalate gel [33] |
| 1,4-Diazabicyclo[2.2.2]octane (DABCO) | Structure-directing agent for supramolecular assemblies | Organic cation template for hybrid compounds [34] |
| Oxalic acid dihydrate (HâCâOâ·2HâO) | Chelating ligand for metal coordination | Precipitation of metal oxalate precursors [33] |
| Citric acid (CâHâOâ) | Gelation agent and metal chelator | Sol-gel synthesis of mixed metal oxides [33] |
| Diethyl hex-2-enedioate | Diethyl Hex-2-enedioate|CAS 21959-75-1|RUO | |
| Phenol--oxotitanium (2/1) | Phenol--oxotitanium (2/1), CAS:20644-86-4, MF:C12H12O3Ti, MW:252.09 g/mol | Chemical Reagent |
The following diagram illustrates the complete experimental workflow from precursor synthesis to final catalytic testing:
The molecular-level processes during supramolecular precursor assembly and thermal decomposition are illustrated below:
Precursor-based synthesis through calcination of inorganic complexes represents a versatile and powerful approach for designing advanced catalytic nanomaterials with tailored properties. The method provides exceptional control over compositional, structural, and morphological characteristics by leveraging molecular-level design in the precursor stage. Recent advances in supramolecular chemistry have further enhanced this paradigm through the development of organic-inorganic hybrid precursors that enable even more precise spatial control over inorganic component organization.
Future developments in this field will likely focus on several key areas: (1) designing increasingly sophisticated supramolecular precursors with multi-functionality; (2) developing milder calcination protocols that preserve delicate nanostructures; (3) integrating computational materials design with experimental synthesis for predictive precursor selection; and (4) scaling laboratory synthesis to industrial production while maintaining nanoscale precision. As research continues to bridge molecular chemistry with materials science, precursor-based calcination routes will remain indispensable for advancing catalytic nanomaterial technology.
The strategic design and functionalization of ligands constitute a cornerstone of modern inorganic chemistry, particularly in the synthesis of supramolecular clusters. The integration of specific ligand typesâpyrazine, carboxylates, and N-heterocyclic carbenes (NHCs)âenables precise control over the electronic properties, structural topology, and functional applications of the resulting metal-organic assemblies. Within the broader context of inorganic cluster supramolecular compound synthesis, ligand substitution and functionalization emerge as critical methodologies for tuning cluster characteristics for specialized applications in catalysis, materials science, and drug development [35] [36].
This technical guide examines the fundamental properties, coordination behavior, and functionalization strategies of these three distinct ligand classes, providing researchers with a comprehensive framework for their implementation in advanced cluster synthesis. The deliberate selection and engineering of these ligands facilitate the construction of complex architectures with tailored functionalities, pushing the boundaries of supramolecular chemistry [37] [38].
Pyrazine belongs to the diazine family, characterized by a six-membered ring containing two nitrogen atoms at para positions. Compared to its monodentate counterpart pyridine, pyrazine exhibits distinct electronic properties that significantly influence its coordination behavior and the resulting metal complexes [37].
Table 1: Comparative Electronic Properties of Nitrogen Heterocycles in Ligand Design
| Heterocycle | pKa of Conjugate Acid | Ï-Acidity (TOTÏD) | Aromaticity (NICS) | Primary Coordination Mode |
|---|---|---|---|---|
| Pyridine | 5.2 | Moderate | -8.9 | Ï-donation via N lone pair |
| Pyrazine | 1.2 | High | -8.5 | Bridging via two N atoms |
| Pyrimidine | 1.3 | Higher | -10.1 | Ï-donation via N lone pair |
| 1,3,5-Triazine | 0.6 | Very high | -12.5 | Ï-donation via N lone pair |
The decreased basicity of pyrazine (pKa = 1.2) compared to pyridine (pKa = 5.2) results from the electron-withdrawing effect of the second nitrogen atom, which stabilizes the lone pair and reduces proton affinity. Conversely, the Ï-accepting ability of pyrazine is enhanced due to stabilization of the Ï* molecular orbitals, facilitating back-bonding from electron-rich metal centers [37]. This electronic configuration makes pyrazine particularly effective as a bridging ligand in cluster compounds, enabling the construction of extended architectures with enhanced electronic communication between metal centers [39].
Ligand substitution kinetics studies on pentacyanoruthenate(II) complexes with pyrazine derivatives reveal a dissociative interchange mechanism, with dissociation rates (kââ) ranging between 2Ã10â»âµ and 10Ã10â»âµ sâ»Â¹ at 25°C. The formation rates from [Ru(CN)â (HâO)]³⻠are typically first order with respect to ligand concentration, with rate constants approximately 1 dm³ molâ»Â¹ sâ»Â¹ for neutral ligands [39].
Carboxylate ligands represent another crucial class in cluster assembly, characterized by their versatile binding modes and ability to form stable bridges between metal centers. The bidentate chelating capability of carboxylates enhances cluster stability through the "chelate effect," while their anionic character modulates the overall charge distribution within the supramolecular architecture [40].
In supramolecular chemistry, carboxylates can be strategically incorporated into more complex ligand systems. For instance, bidentate carboxylate chelating N-heterocyclic carbene ligands have been designed to provide exceptional stability under strong reducing conditions, resisting reductive eliminationâa valuable property for catalytic applications [40]. The synergistic combination of carboxylate and NHC donor functions within a single ligand framework creates multifunctional platforms for advanced cluster design with tailored redox properties and coordination geometries.
N-Heterocyclic Carbenes have emerged as powerful ligands in supramolecular chemistry due to their exceptional Ï-donor capabilities and moderate Ï-acceptor character. These properties enable the formation of strong metal-carbon bonds, resulting in complexes with remarkable stability and unique reactivity profiles [40] [36].
The integration of NHCs into macrocyclic architectures and molecular cages represents a significant advancement in supramolecular chemistry. Poly-NHC precursors serve as versatile building blocks for constructing two-dimensional molecular metallacycles and three-dimensional metallacages with defined cavities and functionalities [36]. These assemblies find applications in molecular recognition, luminescent materials, and catalysis, leveraging the robust metal-carbene bonds to maintain structural integrity under demanding conditions [38] [36].
Functionalization of NHC ligands with specific moieties such as adamantyl groups enables supramolecular interactions with host molecules like cucurbiturils, facilitating non-covalent immobilization strategies. This approach combines the high efficiency of homogeneous catalysts with the practical separation advantages of heterogeneous systems [40].
Table 2: Characteristic Properties and Applications of Ligand Classes in Cluster Synthesis
| Ligand Class | Key Properties | Common Coordination Modes | Cluster Applications |
|---|---|---|---|
| Pyrazine | Moderate Ï-acceptor, low Ï-donor, bridging capability | N,N'-bridging, monodentate | Electronic communication, extended structures |
| Carboxylate | Anionic, strong chelator, versatile binding | Bridging, bidentate chelating, monodentate | Charge balance, structural diversity, stability |
| N-Heterocyclic Carbenes | Strong Ï-donor, moderate Ï-acceptor, tunable sterics | Monodentate, bridging, pincer-type | Stable metal-ligand bonds, catalysis, biomedicine |
The incorporation of adamantyl moieties into NHC ligands enables host-guest chemistry with macrocyclic hosts such as cucurbiturils, facilitating supramolecular immobilization strategies. The following protocol outlines the synthesis of a bidentate carboxylate chelating N-heterocyclic carbene ligand with adamantyl functionality [40]:
Experimental Protocol 1: Synthesis of Adamantyl-Modified Benzimidazolium Salt
Preparation of Dimethyl 1H-Benzo[d]imidazole-5,6-dicarboxylate (3): Suspend 1H-benzo[d]imidazole-5,6-dicarboxylic acid (2) (4.0 g, 19.42 mmol) in 200 mL methanol. Add 4 mL concentrated HâSOâ and reflux the reaction mixture for 16 hours.
Work-up: Evaporate the solvent and neutralize the viscous liquid with 6M KâCOâ until precipitate formation ceases. Add 200 mL DCM, extract, and evaporate the organic layer to dryness to obtain a white solid (yield: 4.03 g, 89%).
N-Alkylation: Dissolve compound (3) (4.00 g, 17.10 mmol) and p-tolyboronic acid (3.34 g, 25.65 mmol) in 190 mL distilled methanol. Add Cu(NOâ)â (0.824 g, 3.42 mmol) and TMEDA (0.255 mL, 1.71 mmol). Stir at room temperature for 16 hours under oxygen atmosphere.
Purification: Remove solvent and dissolve the residue in 100 mL DCM. Extract with water (3 Ã 50 mL). Purify the crude product by flash column chromatography (SiOâ, 90:10 CHâClâ/MeOH) to obtain dimethyl 1-p-tolyl-1H-benzo[d]imidazole-5,6-dicarboxylate (4) as a light yellow solid (yield: 5.36 g, 96%).
Adamantyl Functionalization: Further reactions introduce the adamantyl moiety to complete the NHC ligand precursor. Characterization includes ¹H-NMR, ¹³C-NMR, FTIR, single crystal X-ray crystallography, and elemental analysis [40].
The adamantyl-modified NHC ligand can be non-covalently immobilized through host-guest interactions with cucurbit[7]uril (CB7):
Titration Method: Perform ¹H-NMR titration of the NHC ligand (8) with CB7 in appropriate deuterated solvent.
Binding Assessment: Monitor chemical shift changes in the adamantyl proton signals, indicating encapsulation within the CB7 cavity.
Immobilization Confirmation: Characterize the resulting supramolecular complex by NMR spectroscopy, observing distinctive upfield shifts for adamantyl protons due to the shielding environment of the host macrocycle [40].
The construction of supramolecular clusters from poly-NHC ligand precursors employs several strategic approaches, each offering distinct advantages for specific architectural targets [36]:
1. One-Pot Synthesis: Direct reaction of poly-NHC precursors with metal salts under controlled conditions, suitable for thermodynamically favored symmetric structures.
2. Supramolecular Transmetalation: Utilizes labile metal-ligand bonds in precursor complexes to form more stable structures via metal exchange.
3. Stepwise Synthesis: Sequential construction of complex architectures through controlled addition of building blocks, allowing for precise incorporation of functional groups.
4. Improved One-Pot Strategy with Self-Sorting: Leverages the selective recognition between specific ligand and metal components to form complex architectures from a mixture of precursors.
5. Subtle Variation Strategy: Minor modifications to poly-NHC precursor structures to direct the formation of different supramolecular architectures from similar building blocks.
These methodologies enable the precise construction of two-dimensional molecular metallacycles and three-dimensional metallacages with defined shapes, sizes, and functionalities, advancing applications in catalysis, molecular encapsulation, and materials science [36].
Diagram 1: Strategic workflow for supramolecular cluster assembly through rational ligand design and functionalization, highlighting key decision points in ligand selection and assembly methodology.
Comprehensive characterization of functionalized ligands and their supramolecular clusters necessitates a multidisciplinary analytical approach. Advanced spectroscopic and structural elucidation techniques provide insights into molecular architecture, bonding, and host-guest interactions.
Nuclear Magnetic Resonance (NMR) Spectroscopy: Multinuclear NMR (¹H, ¹³C, ¹¹B, etc.) serves as a primary tool for characterizing ligand functionalization and host-guest interactions. For adamantyl-modified NHC ligands, ¹H-NMR titration with cucurbituril hosts reveals complex formation through characteristic chemical shift changes of adamantyl protons [40]. Diffusion-ordered spectroscopy (DOSY) can further elucidate the size and stability of supramolecular assemblies in solution.
Vibrational Spectroscopy: Infrared spectroscopy identifies characteristic functional group vibrations, including C=O stretches of carboxylates (1650-1550 cmâ»Â¹) and C=N stretches of N-heterocyclic systems. Raman spectroscopy complements IR data, particularly for symmetric vibrations and metal-cluster bonds.
Electronic Spectroscopy: UV-Vis absorption spectroscopy monitors charge-transfer transitions and electronic communication between metal centers in bridged clusters. Emission spectroscopy proves valuable for luminescent clusters, with modifications to ligand structures directly influencing photophysical properties [36].
X-ray Crystallography: Single-crystal X-ray diffraction remains the definitive method for determining molecular and supramolecular structures. It provides precise bond lengths, angles, and coordination geometries, essential for understanding the structural consequences of ligand functionalization [40] [35].
For cluster compounds, crystallography reveals:
Elemental Analysis: Combustion analysis verifies elemental composition, confirming successful ligand functionalization and purity of synthetic products.
The strategic functionalization of ligands with pyrazine, carboxylate, and NHC motifs enables the construction of supramolecular clusters with tailored properties for advanced applications.
Supramolecular clusters incorporating functionalized ligands serve as versatile catalysts for various transformations. NHC-based metallacages provide confined microenvironments that enhance substrate preorganization and reaction selectivity [36]. The integration of carboxylate chelating groups in NHC ligands confers resistance against reductive elimination under strong reducing conditions, expanding their utility in catalytic hydrogenation and related processes [40].
Non-covalent immobilization of adamantyl-functionalized NHC catalysts onto solid supports via host-guest interactions with surface-bound cucurbiturils or cyclodextrins combines the advantages of homogeneous catalysis (high activity and selectivity) with heterogeneous recovery and recyclability [40].
NHC-based supramolecular assemblies show promising biomedical applications, particularly in anticancer therapy. Silver(I) and gold(I) NHC complexes exhibit potent cytotoxic activity against various cancer cell lines, with the supramolecular architecture influencing cellular uptake and biological activity [36]. The encapsulation of drug molecules within metallacages formed by poly-NHC ligands can modify pharmacokinetic profiles and enhance therapeutic efficacy through controlled release mechanisms.
Luminescent supramolecular clusters incorporating functionalized ligands find applications in sensing, imaging, and optoelectronics. The modulation of electronic properties through pyrazine bridging or carboxylate functionalization directly influences emission characteristics, enabling the design of materials with tailored photophysical properties [38] [36].
Table 3: Research Reagent Solutions for Ligand Functionalization and Cluster Synthesis
| Reagent/Category | Function/Purpose | Application Examples |
|---|---|---|
| Pentacyanoruthenate(II) | Metal precursor for kinetic studies | Investigating ligand substitution rates of pyrazine derivatives [39] |
| Cucurbit[7]uril (CB7) | Macrocyclic host for supramolecular chemistry | Non-covalent immobilization of adamantyl-modified NHC ligands [40] |
| Poly-NHC Precursors | Building blocks for supramolecular assemblies | Construction of 2D metallacycles and 3D metallacages [36] |
| Adamantyl Derivatives | Supramolecular recognition elements | Introducing host-guest binding capabilities to NHC ligands [40] |
| Bidentate Carboxylate Chelators | Stable coordinating units with charge balance | Enhancing cluster stability under reducing conditions [40] |
Ligand substitution and functionalization with pyrazine, carboxylate, and N-heterocyclic carbene motifs provide powerful strategies for tuning the properties of supramolecular clusters. The systematic modification of electronic characteristics, binding modes, and supramolecular recognition elements enables precise control over cluster architecture, stability, and function.
The integration of these ligand classes within a unified supramolecular approach facilitates the development of advanced materials with applications spanning catalysis, biomedicine, and materials science. As synthetic methodologies advance and our understanding of structure-property relationships deepens, the strategic design of functionalized ligands will continue to drive innovation in inorganic cluster supramolecular chemistry.
Future directions will likely focus on increasing structural complexity through hierarchical assembly, enhancing functionality through multi-component ligand systems, and expanding applications in targeted drug delivery and alternative energy technologies.
The synthesis of inorganic cluster supramolecular compounds represents a frontier in materials science, enabling the design of complex architectures with tailored properties for applications ranging from catalysis to drug delivery. This field leverages the principles of supramolecular chemistry, where non-covalent interactionsâprimarily hydrogen bonding and Ï-Ï stackingâgovern the self-assembly of discrete molecular components into extended, functional frameworks [41]. Within the context of a broader thesis on inorganic cluster supramolecular compound synthesis, this guide details the core interactions, characterization methodologies, and experimental protocols essential for constructing these sophisticated structures. The dynamic and reversible nature of these interactions facilitates the formation of systems that are both structurally robust and adaptively responsive to external stimuli [42].
The synergy between hydrogen bonding and Ï-Ï stacking is particularly critical. Hydrogen bonds provide directionality and specificity, while Ï-Ï interactions between aromatic systems confer structural stability and facilitate long-range order through layer-by-layer stacking [43] [44]. Incorporating pre-formed inorganic clusters, such as the GeâSâââ´â» cluster, as building blocks introduces well-defined nanoscopic functionality and can template the overall framework topology [31]. The following sections provide a technical guide for researchers and scientists aiming to master the synthesis and characterization of these advanced materials.
Hydrogen bonding is a directional, moderately strong non-covalent interaction (typically ranging from 4 to 21 kJ molâ»Â¹) that is fundamental to supramolecular assembly [42]. It occurs between a hydrogen atom bound to an electronegative donor (e.g., N-H, O-H) and an electronegative acceptor (e.g., N, O, S).
In the context of inorganic cluster frameworks, a variety of hydrogen bonds can be employed to form multi-dimensional networks. These include:
For instance, in the structure of (Hâbipy)âGeâSâââ (bipy)â 7HâO, the monoprotonated 4,4'-bipyridin-1-ium cations form supramolecular chains via N-Hâ â â N hydrogen bonds. These chains are further connected into a three-dimensional architecture through additional interactions involving water molecules and counter-ions [31] [44].
Ï-Ï stacking refers to attractive, non-covalent interactions between the Ï-orbitals of aromatic rings. The strength of this interaction is influenced by factors such as the electron density and substitution patterns of the aromatic systems involved [45]. Key modes include:
These interactions are crucial for stabilizing layered structures and creating extended supramolecular architectures. The planar pyrene core in HâTCPy (1,3,6,8-tetracarboxy pyrene), for example, provides a strong layer-by-layer Ï-Ï stacking interaction that reinforces the hydrogen-bonded organic framework (HOF) over a wide pH range (1-11) [43].
The combination of hydrogen bonding and Ï-Ï stacking is a powerful strategy for building stable, extended frameworks. Hydrogen bonds often define the primary assembly and provide specificity, while Ï-Ï interactions consolidate the structure by stabilizing the stacking of layers or chains. This synergy is exemplified in a Ï-stacked supramolecular cage (Ï-MX-cage), which is assembled from 16 [MXL]⺠ions via 18 intermolecular Ï-stacking interactions. The resulting tetrahedral cage presents 48 exterior N-H hydrogen bond donors, which are available for further hierarchical assembly with guest molecules through hydrogen bonding [46].
Table 1: Key Non-Covalent Interactions in Supramolecular Framework Synthesis
| Interaction Type | Typical Energy (kJ molâ»Â¹) | Role in Framework Assembly | Common Functional Groups/Moieties |
|---|---|---|---|
| Hydrogen Bonding | 4 - 21 [42] | Provides directionality and specificity; forms primary network topology. | N-H, O-H (donors); N, O, S (acceptors) [31] |
| Ï-Ï Stacking | 5 - 20 (estimated) | Stabilizes layered structures; enables long-range order and electronic communication. | Aromatic rings (e.g., pyrene, bipyridine, benzene derivatives) [43] [45] |
| Cation-Ï | 5 - 80 [45] | Binds cationic species; important for biological recognition and sensing. | Aromatic rings (electron-rich), cationic groups (e.g., -NRââº) [45] |
| Anion-Ï | Variable (weaker) | Binds anionic species; used in sensing and environmental remediation. | Electron-deficient aromatic rings (e.g., tetracyanoquinodimethane) [45] |
Confirming the structure and stability of supramolecular frameworks requires a suite of analytical techniques.
Table 2: Key Characterization Methods and Their Information Output
| Technique | Key Information | Application Example |
|---|---|---|
| SCXRD | 3D atomic structure, interaction geometries (H-bond distances, Ï-stacking distances) | Determining the head-to-tail N-Hâ â â N hydrogen-bonded chains in 4,4'-bipyridin-1-ium compounds [44]. |
| PXRD | Bulk crystallinity, phase purity | Confirming that a synthesized HRP@HOF-100 biocomposite matches the crystalline pattern of standard HOF-100 [43]. |
| Cryo-EM | Microstructure, spatial distribution of components, pore channels | Visualizing the long-range ordered pore channels in a HOF hybrid biocatalyst [43]. |
| TGA | Thermal stability, solvent content, decomposition profile | Observing weight loss at 200â350°C due to pyrolysis of encapsulated proteins in HRP@HOF-100 [43]. |
| Gas Physisorption | Surface area, pore volume, pore size distribution | Calculating a BET surface area of 571.2 m²/g for HRP@HOF-100, indicating microporosity and partial pore occupation [43]. |
| FTIR Spectroscopy | Functional group identification, evidence of H-bonding | Detecting the amide I band of an enzyme in a HOF composite, confirming successful encapsulation [43]. |
The following protocol, adapted from the synthesis of (Hâbipy)âGeâSâââ
(bipy)â
7HâO, details a hydrothermal method for constructing a hybrid framework incorporating a germanium sulfide cluster [31].
Objective: To synthesize a crystalline supramolecular network containing GeâSâââ´â» clusters linked via hydrogen bonding and Ï-Ï stacking of 4,4'-bipyridine ligands.
Materials:
Procedure:
Key Considerations:
This protocol describes a green, aqueous-based method for encapsulating enzymes within a hydrogen-bonded organic framework, using HâTCPy as a building block [43].
Objective: To in-situ encapsulate horseradish peroxidase (HRP) within a HOF-100 scaffold under mild, almost organic solvent-free conditions.
Materials:
Procedure:
Key Considerations:
Table 3: Key Reagents for Supramolecular Framework Synthesis
| Reagent / Material | Function in Synthesis | Example Use Case |
|---|---|---|
| TMAâGeâSââ | Source of pre-formed [GeâSââ]â´â» tetrahedral cluster; inorganic building block. | Core component in inorganic-organic hybrid frameworks [31]. |
| 4,4'-bipyridine (bipy) | Rigid, linear organic linker; can be protonated to form hydrogen bonds. | Forms supramolecular chains via N-Hâ â â N H-bonds and provides Ï-surfaces for stacking [31] [44]. |
| HâTCPy | Câ-symmetric pyrene tecton with four formic acid arms; HOF building block. | Self-assembles via H-bonds and Ï-Ï stacking to form HOF-100 for enzyme encapsulation [43]. |
| Tris(2-benzimidazolylmethyl)amine (L) | Multidentate organic ligand for metal coordination. | Used in the assembly of Ï-stacked supramolecular cages with transition metals [46]. |
| Enzymes (e.g., HRP) | Functional biological guest molecules for encapsulation. | Creates hybrid biocatalysts for biosensing and selective catalysis [43]. |
| Benzene, 1-butyl-4-ethyl | Benzene, 1-butyl-4-ethyl, CAS:15181-08-5, MF:C12H18, MW:162.27 g/mol | Chemical Reagent |
| Chloro(isopropyl)silane | Chloro(isopropyl)silane, MF:C3H7ClSi, MW:106.62 g/mol | Chemical Reagent |
The following diagram illustrates the decision-making and experimental workflow for designing and synthesizing supramolecular frameworks via H-bonding and Ï-Ï stacking.
The strategic combination of hydrogen bonding and Ï-Ï stacking provides a powerful and versatile methodology for constructing extended supramolecular frameworks with inorganic clusters. The protocols and characterizations detailed in this guide provide a foundation for the rational design of next-generation materials. As the field progresses, the integration of computational design, the exploration of new dynamic covalent chemistry, and a relentless focus on functional performance will undoubtedly unlock new frontiers in catalysis, sensing, and precision medicine.
Inorganic cluster supramolecular compounds (ICSs) represent an emerging class of materials in anticancer research, formed through the precise organization of metal clusters via supramolecular interactions or templating effects. Supramolecular chemistry, defined as 'chemistry beyond the molecule,' focuses on molecular assemblies held together by non-covalent interactions, enabling the construction of highly sophisticated systems with dynamic, reversible, and stimulus-responsive properties [47] [48]. The synthesis of ICSs typically involves reactions between organic structure-directing agents and inorganic metal salts, yielding hybrid architectures where inorganic clusters are spatially organized by organic cations through weak electrostatic forces and other non-covalent interactions [34].
This review examines the potential of ICSs as anticancer agents, focusing on their unique mechanisms of action and pathways to achieving selective toxicity against cancer cells. Unlike conventional chemotherapy that often lacks specificity, ICSs offer the possibility of targeted cancer cell destruction through novel biological interference, enhanced drug delivery capabilities, and activation by distinctive tumor microenvironment stimuli. The modular nature of ICS synthesis allows for systematic tuning of their physicochemical properties, enabling researchers to optimize these compounds for specific anticancer applications including direct cytotoxicity, drug encapsulation, and combination therapy regimens.
The rational design of ICSs for anticancer applications relies on strategic manipulation of their core components: the inorganic cluster, the organic template or ligand, and their supramolecular interaction interfaces. The inorganic cluster (e.g., tetrachloroaurate, cadmium iodide, copper iodide, or cobalt chloride) provides fundamental photophysical, redox, and catalytic properties that can be harnessed for therapeutic effects [49] [34]. The organic structure-directing component (e.g., cyclodextrins, pillararenes, cucurbiturils, or custom-synthesized cationic ligands) controls assembly geometry, confers biocompatibility, and introduces targeting capabilities [34] [48]. The supramolecular interactions (electrostatic forces, host-guest complexation, hydrophobic effects, hydrogen bonding) enable dynamic, responsive behaviors that can be exploited for triggered drug release or activation within tumor environments [47] [48].
Crystal engineering of these materials reveals diverse structural motifs with potential biomedical relevance. For instance, compounds like {[L][HgI4]} (1) and {[L][CdI4]} (2) exhibit mononuclear anionic structures where metal ions adopt distorted tetrahedral configurations, with bond lengths ranging 2.763-2.808 à for Hg-I and 2.767-2.791 à for Cd-I [34]. These precise structural parameters influence stability, reactivity, and eventual biological interactions. More complex architectures such as {[(L)(Cu2I3)]·[CuI2]CH3CN}n (3) form extended 1D chain structures with rectangular pore-like geometries (15.979 à 16.838 à ) capable of guest molecule encapsulation [34].
Table 1: Essential research reagents for ICS synthesis and evaluation
| Reagent Category | Specific Examples | Function in ICS Development |
|---|---|---|
| Metal Precursors | Tetrachloroaurate, Cadmium iodide, Copper iodide, Cobalt chloride | Inorganic cluster formation with specific coordination geometries and electronic properties |
| Organic Templates | Cyclodextrins (α, β, γ), Pillararenes, Cucurbiturils, DABCO derivatives | Structure direction, charge balance, and modification of ICS architecture |
| Surface Modifiers | Polyethylene glycol (PEG), Thiols, Amines, Phosphines | Biocompatibility enhancement, RES evasion, circulation time extension |
| Targeting Moieties | Peptides, Antibodies, Aptamers | Specific recognition of cancer cell surface markers for selective accumulation |
| Characterization Tools | Dynamic Light Scattering, FTIR Spectroscopy, TEM, X-ray Diffraction | Size distribution analysis, surface functionalization verification, structural elucidation |
ICSs can induce direct cancer cell death through several supramolecularly-enhanced mechanisms. Gold nanoparticle-based ICSs demonstrate particularly promising cytotoxic properties when properly functionalized. In one mechanistic study, ramucirumab-conjugated gold nanoparticles induced direct cytotoxic effects specifically in gastric cancer cells, while neither the antibody nor gold nanoparticles alone showed significant efficacy even at high concentrations [50]. Proteomic and transcriptomic analyses revealed this direct cytotoxicity was predominantly mediated by antibody-enhanced phagocytosis, with high affinity immunoglobulin gamma Fc receptor I showing differential up-regulation in cancer cells treated with the supramolecular construct [50].
The localized surface plasmon resonance (LSPR) characteristic of gold-based ICSs provides another direct anticancer mechanism. When stimulated with near-infrared light (650-900 nm), these particles generate localized hyperthermia through the surface plasmon resonance effect [51]. This photothermal effect enables highly selective tumor ablation, as normal tissues remain relatively transparent to NIR wavelengths, while ICSs accumulated in tumor tissue efficiently convert light energy to heat [51]. The tunability of LSPR properties through adjustments to ICS size, shape, and aspect ratio allows optimization for specific therapeutic applications.
ICSs function as advanced drug delivery platforms through host-guest interactions that encapsulate anticancer drugs within their supramolecular architectures. Cyclodextrin-based ICSs have demonstrated particular utility in this context, as their hydrophobic cavities can entrap appropriately sized chemotherapeutic agents (e.g., doxorubicin, paclitaxel), while their hydrophilic exteriors improve water solubility and bioavailability [48]. This encapsulation approach significantly reduces systemic toxicity while maintaining therapeutic efficacy at tumor sites.
Stimuli-responsive release mechanisms represent a key advantage of ICS-based drug delivery. These systems can be engineered to release their payload in response to specific tumor microenvironment cues, including:
These controlled release mechanisms enhance selective toxicity by maximizing drug exposure to malignant cells while minimizing off-target effects.
ICSs provide ideal platforms for implementing combination therapy approaches that target multiple cancer pathways simultaneously. Supramolecular materials enable precise spatial organization of therapeutic components, facilitating synergistic interactions between treatment modalities [48]. Notable combination strategies include:
Chemotherapy-Photothermal Therapy: ICSs can concurrently deliver conventional chemotherapeutic agents while serving as photothermal transducers. This approach was demonstrated with gold nanorod-based systems conjugated with doxorubicin and targeting antibodies, enabling combined chemotherapy, photothermal ablation, and molecular targeting in a single construct [50].
Chemotherapy-Gene Therapy: Supramolecular polymers constructed through host-guest interactions between camptothecin-conjugated β-cyclodextrin and therapeutic oligonucleotides have shown promise in co-delivering chemotherapeutic and gene-modulating agents [48]. The disulfide linkages in these systems facilitate glutathione-responsive release in cancer cells, enhancing selective toxicity.
Dual-Drug Chemotherapy: Self-assembled supramolecular nanoparticles (SSNPs) have been developed for co-delivering combination chemotherapies (e.g., doxorubicin and docetaxel). These systems employ β-cyclodextrin polymers as structural frameworks that host two anticancer drugs simultaneously, with surface modification by targeting aptamers to enhance tumor selectivity [48].
ICS Synthesis via Solvent Volatilization Method:
Functionalization with Targeting Moieties:
Cytotoxicity Profiling:
Cellular Uptake and Localization:
Table 2: Comparative efficacy of supramolecular anticancer formulations
| ICS Formulation | Cancer Model | Key Performance Metrics | Selective Toxicity Index |
|---|---|---|---|
| AuNR-PEG-Ab-DOX | Gastric cancer (SNU-5) | 3Ã higher cellular uptake vs. free Ab; Significant cytotoxicity only in combination [50] | >10Ã tumor vs. normal tissue accumulation [50] |
| β-CD-DOX/DTX SSNPs | Multiple cancer types | Enhanced cytotoxicity to target cells; Positive synergistic effect [48] | Aptamer-mediated targeting to cancer cells [48] |
| CPT-CD Supramolecular Polymer | Various malignancies | Glutathione-responsive drug release; Improved therapeutic efficacy [48] | Selective release in high-GSH cancer environment [48] |
| POSS-(sulfobetaine)7/CD-PLLA | Tumor models | Effective tumor inhibition; Sustained drug release [48] | Reduced toxicity to normal cells [48] |
| pH-sensitive β-CD-g-PDMAEMA@Azo-PCL | Experimental cancers | Accelerated drug release at pH 4 vs. pH 7 [48] | 5.2à faster release in acidic tumor environment [48] |
Inorganic cluster supramolecular compounds represent a promising frontier in targeted cancer therapy, offering unique mechanisms of action that enhance selective toxicity against malignant cells. The supramolecular design principles underlying ICS development enable precise control over their interactions with biological systems, facilitating cancer-specific accumulation, activation, and therapeutic effects. Through direct cytotoxicity, enhanced drug delivery, and combination therapy approaches, ICSs address fundamental limitations of conventional chemotherapy.
The continued advancement of ICS anticancer agents will require interdisciplinary collaboration across supramolecular chemistry, materials science, and oncology. Future research directions should focus on optimizing supramolecular design for enhanced in vivo stability, tumor targeting precision, and responsiveness to tumor-specific stimuli. As understanding of supramolecular behavior in biological systems deepens, ICS-based approaches hold significant potential for clinical translation, ultimately contributing to more effective and selective cancer treatments with reduced adverse effects.
Inorganic Cluster Supramolecules (ICSs) represent a class of compounds where inorganic metal clusters are integrated into larger, organized structures through supramolecular chemistry, leveraging non-covalent interactions. Within the broader context of inorganic cluster supramolecular compound synthesis, these materials are engineered to possess well-defined cavities and binding sites, making them ideal candidates for host-guest chemistry. This property is directly exploitable in pharmaceutical sciences for the targeted delivery of therapeutic agents. The fundamental principle involves the encapsulation of drug molecules (guests) within the scaffolds of these inorganic clusters (hosts), forming stable host-guest complexes that can protect the drug and control its release under specific physiological conditions [52]. This approach aligns with advanced drug delivery strategies that seek to improve therapeutic efficacy while minimizing side effects.
The synthesis of these compounds often results in unique hybrid architectures. For instance, research has detailed the creation of a unique mixed-valence CuII/CuI organicâinorganic hybrid supramolecular cluster, {[Cu(DMSO)5][Cu4I6(DMSO)]}n, which demonstrates a three-dimensional supramolecular framework stabilized by hydrogen bonds [53]. Another study reports on inorganic-organic hybrid assemblies incorporating 12-silicotungstic acid heteropolyoxometalate and trinuclear lanthanide clusters [54]. Such structures highlight the synthetic versatility in creating multifunctional materials where the inorganic cluster provides structural integrity and potential functionality (e.g., magnetism, luminescence), while the supramolecular architecture enables guest encapsulation.
The encapsulation of therapeutics within ICSs is governed by host-guest chemistry, a subset of supramolecular chemistry focused on the formation of complexes between a host molecule and a guest molecule. Macrocyclic hosts like cyclodextrins, calixarenes, cucurbiturils, and pillararenes are well-known in theranostics for their ability to form inclusion complexes with various drug molecules [52]. The binding affinity in these complexes is quantified by the association constant (Ka), a critical parameter that determines the stability of the host-guest system under physiological conditions [55]. The dynamic and reversible nature of the non-covalent interactions (e.g., hydrogen bonding, halogen bonding, van der Waals forces, hydrophobic effects) that underpin host-guest complexation is key to designing responsive drug delivery systems. These interactions can be modulated by the local microenvironment, allowing for targeted drug release at disease sites [52].
Accurately measuring the strength of host-guest interactions is vital for developing effective ICS-based delivery systems. The association constant (Ka) provides a quantitative measure of this binding affinity. Traditional methods for determining Ka include NMR, UV-vis, and fluorescence spectroscopy, as well as isothermal titration calorimetry (ITC) [55]. However, recent advances have introduced in situ Fourier-Transform Infrared (FTIR) spectroscopy as a powerful alternative. This technique probes specific bond vibrations, offering a fast timescale that can capture dynamic conformational changes during guest binding that might be averaged out in NMR measurements [55]. A typical application involves monitoring the shift in carbon-deuterium (C-D) stretching frequencies within the IR-transparent window (1800â2500 cmâ»Â¹) upon anion binding to an imidazolium-based host, enabling the calculation of Ka through global fitting of the spectral data [55].
Table 1: Techniques for Measuring Host-Guest Association Constants (Ka)
| Technique | Principle | Advantages | Limitations |
|---|---|---|---|
| NMR Spectroscopy | Chemical shift changes upon binding. | Provides structural information. | Time-averaged snapshot; may require deuterated solvents. |
| UV-vis/Fluorescence | Changes in chromophore/fluorophore properties. | High sensitivity for conjugated systems. | Requires the presence of a suitable chromophore/fluorophore. |
| Isothermal Titration Calorimetry (ITC) | Direct measurement of heat change during binding. | Provides full thermodynamic profile (ÎH, ÎS). | Requires relatively high concentrations of samples. |
| In situ FTIR Spectroscopy | Shifts in vibrational frequencies of specific bonds. | Fast timescale; minimal sample prep; no need for deuterated solvents. | Lower molar absorptivity for some vibrations (e.g., C-D). |
The synthesis of ICSs and their subsequent loading with therapeutic agents necessitates rigorous characterization to confirm structure, purity, and successful encapsulation. Single-crystal and powder X-ray diffraction are paramount for determining the precise atomic arrangement and confirming the formation of the supramolecular framework [53]. Elemental analysis and infrared (IR) spectroscopy further verify the chemical composition and identify functional groups involved in host-guest interactions [53] [55]. Thermogravimetric analysis (TGA) assesses the thermal stability of the ICS, which is crucial for understanding its behavior during processing and storage [53].
Beyond structural confirmation, evaluating the functional performance of ICS-based delivery systems is essential. Luminescence spectroscopy can be employed to study the photophysical properties of the cluster and to develop sensing applications, such as the detection of nitroaromatic compounds like 2,4,6-trinitrophenol (TNP) [53]. More critically for drug delivery, encapsulation efficiency (EE) and drug release kinetics are key metrics. While specific EE data for ICSs can be limited, studies on nanocarriers like solid lipid nanoparticles (SLNs) report encapsulation efficiencies for hydrophobic vitamins in the range of 27â45% [56]. The release kinetics of drugs from polymeric carriers often follow a power-law expression (Mt/Mâ = ktâ¿), where 'n' is the release exponent indicating the transport mechanism (e.g., Fickian diffusion, Case-II transport) [57]. These principles are directly applicable to characterizing drug release from ICSs.
Table 2: Key Analytical Techniques for ICS Characterization
| Technique | Primary Function | Application Example |
|---|---|---|
| Single-Crystal X-ray Diffraction | Determine atomic-level 3D structure. | Confirming the wavelike anionic chain in {[Cu(DMSO)5][Cu4I6(DMSO)]}n [53]. |
| FTIR Spectroscopy | Identify functional groups and probe binding interactions. | Monitoring C-D bond redshift to calculate Ka for host-guest complexation [55]. |
| Thermogravimetric Analysis (TGA) | Measure thermal stability and decomposition profile. | Determining the temperature at which an ICS lattice loses solvent molecules [53]. |
| Luminescence Spectroscopy | Investigate optical properties and sensing capabilities. | Probing the quenching of ICS luminescence by 2,4,6-trinitrophenol for sensing [53]. |
Protocol Objective: Hydrothermal synthesis of {[Cu(DMSO)5][Cu4I6(DMSO)]}n, a unique mixed-valence CuII/CuI supramolecular cluster [53].
Materials:
Methodology:
Protocol Objective: Determine the association constant (Ka) for an anion binding to a deuterated imidazolium host (D-IPr·PF6) using in situ FTIR spectroscopy [55].
Materials:
Methodology:
Diagram 1: Workflow for ICS Synthesis and Binding Analysis
The release of encapsulated therapeutics from ICSs can be engineered to respond to specific stimuli present in the disease microenvironment, a hallmark of intelligent drug delivery systems. The dynamic non-covalent bonds of the host-guest complex can be disrupted by changes in pH, redox potential, or the presence of specific enzymes, leading to controlled drug release [52]. Furthermore, the physical structure of the carrier plays a critical role. Release can be governed by diffusion, where the drug diffuses through the pores of the ICS matrix, or by a degradation mechanism, where the breakdown of the ICS scaffold triggers drug release [57]. For instance, in polymeric systems, drug release from a non-degradable, fully swollen matrix is often described by Fickian diffusion models, whereas in biodegradable systems, drug release is coupled to the erosion rate of the polymer backbone [57]. The goal is to achieve a spatiotemporal release profile that maintains the drug concentration within the therapeutic window for the desired duration.
Diagram 2: Stimuli-Responsive Drug Release Mechanism
The synthesis and application of ICSs for drug delivery require a suite of specialized reagents and materials. The following table details key components for research in this field, drawing from the protocols and systems discussed.
Table 3: Essential Research Reagents for ICS Drug Delivery Development
| Reagent/Material | Function | Example Application |
|---|---|---|
| Metal Salts (e.g., CuI) | Inorganic precursor for cluster formation. | Core component in the synthesis of copper-halide clusters [53]. |
| Organic Ligands (e.g., DMSO, pydc-OH) | Coordinate to metal centers and bridge clusters; form the organic-inorganic hybrid structure. | DMSO coordinates Cu ions; 4-hydroxypyridine-2,6-dicarboxylic acid (pydc-OH) links lanthanide clusters [53] [54]. |
| Deuterated Host Molecules | Enable probing of specific host-guest interactions via FTIR spectroscopy. | C-D bond in imidazolium host allows monitoring of anion binding [55]. |
| Tetra-n-butylammonium (TBA) Salts | Provide soluble, organically-compatible anionic guests for binding studies. | Used to titrate Clâ», Brâ», and Iâ» into host solutions for Ka determination [55]. |
| Polyoxometalates (POMs) | Act as large, anionic inorganic building blocks with diverse structures and properties. | [SiWââOââ]â´â» Keggin ion used to construct hybrid assemblies with lanthanides [54]. |
| Anhydrous, Spectroscopic-Grade Solvents | Ensure reaction integrity and high-quality spectral data without interference. | Essential for accurate in situ FTIR titrations in organic solvents like acetone [55]. |
| Lanthanum--nickel (2/7) | Lanthanum--nickel (2/7), CAS:12532-78-4, MF:La2Ni7, MW:688.66 g/mol | Chemical Reagent |
| Rhodium--vanadium (1/3) | Rhodium--vanadium (1/3), CAS:12210-74-1, MF:RhV3, MW:255.730 g/mol | Chemical Reagent |
The convergence of diagnostic and therapeutic functions into a single agent, known as theranostics, represents a transformative approach in precision medicine. Multimodal theranostic platforms specifically integrate multiple imaging modalities with targeted therapeutic interventions within unified nanosystems. These platforms are fundamentally enabled by supramolecular chemistryâthe study of molecular systems held together by non-covalent interactionsâwhich provides the synthetic framework for constructing complex, functional architectures from discrete molecular components [10] [58]. The philosophical roots of supramolecular chemistry trace back to Hermann Emil Fischer's 1894 "lock and key" hypothesis of enzyme-substrate interactions, establishing the foundational principles of molecular recognition that underpin modern theranostic agent design [59] [10].
The structural and functional paradigm for these platforms increasingly derives from inorganic cluster supramolecular compounds, which offer unparalleled opportunities for synthesizing systems with tailored magnetic, optical, and electronic properties [59] [34]. These coordination-based assemblies demonstrate that the instruction set for creating large complex structures is contained within their constituent components, an concept Jean-Marie Lehn aptly described as an 'information science' [58]. By exploiting weak, reversible non-covalent interactionsâincluding hydrogen bonding, metal coordination, hydrophobic forces, van der Waals forces, Ï-Ï interactions, and electrostatic effectsâresearchers can engineer systems that self-assemble into precise architectures with multiple integrated functionalities [60] [10]. This approach allows for the creation of sophisticated nanoscale devices capable of targeting specific tissues, reporting their location through multiple imaging techniques, and delivering therapeutic payloads in response to biological triggers.
The rational design of inorganic cluster supramolecular compounds hinges on controlling the interplay of several non-covalent forces that direct self-assembly processes. These interactions, while individually weak, collectively determine the structural integrity and functional properties of the resulting assemblies [10]. Metal coordination provides primary directional bonds between organic ligands and metal ions, forming the foundational architecture of the complex [60]. Simultaneously, hydrogen bonding offers secondary structural stabilization, with functional groups like pyrazinones, pyrimidinones, and quinazolinones consistently delivering required synthetic vectors in coordination polymers [60]. Additional Ï-Ï stacking interactions between aromatic systems contribute to molecular recognition and stability, while electrostatic interactions and hydrophobic effects further guide the self-assembly of discrete building blocks into predictable topologies [10] [58].
The robustness of these assembly processes is demonstrated by their insensitivity to potential disruptors. Recent studies with CdII coordination polymers show that directional intermolecular NâHâ¯O hydrogen bonds between ligands on adjacent complex building blocks reliably drive assembly into predictable architectures, regardless of different steric and geometric demands from counterions [60]. Furthermore, these structural outcomes remain consistent whether solids are prepared from solution or through liquid-assisted grinding, emphasizing the reliability of these supramolecular synthetic protocols [60].
A particularly versatile class of building blocks for supramolecular theranostics centers on azabenzene-based ligands, which contain six-membered aromatic rings with one or more nitrogen atoms in the heterocyclic ring [59]. These heterocyclic azines are categorized based on the number (n) of nitrogen atoms present: n = 1 (pyridine); n = 2 (diazines: pyrimidine, pyridazine, pyrazine); n = 3 (1,3,5-triazine); and n = 4 (1,2,4,5-tetrazine) [59]. These ligands are particularly valuable because they function as weak Ï donors and strong Ï acceptors, making them electron-deficient and capable of forming stable coordination complexes with various metal ions [59].
The self-assembly of these ligands with metal ions can produce sophisticated architectures such as [n à n] metallogrid complexes, where 2n ligands arrange like grid poles with n² metal ions situated at the intersections [59]. These metallogrids demonstrate how supramolecular chemistry enables the bottom-up fabrication of nanoscale structures with precisely controlled geometries. The iron and cobalt complexes of these ligands frequently exhibit intriguing magnetic properties and sometimes display spin crossover (SCO) phenomenaâtransition between low spin and high spin states in response to temperature, pressure, or lightâwhich can be exploited for stimulus-responsive applications in data storage, molecular switches, and optical displays [59].
Table 1: Azabenzene-Based Ligands in Supramolecular Assembly
| Ligand Class | Number of N Atoms | Examples | Key Properties | Applications in Supramolecular Assembly |
|---|---|---|---|---|
| Pyridines | 1 | Pyridine | Weak Ï donor, strong Ï acceptor | Basic building block for coordination complexes |
| Diazines | 2 | Pyrimidine, Pyridazine, Pyrazine | Electron-deficient | Formation of metallogrid complexes with transition metals |
| Triazines | 3 | 1,3,5-triazine | Multiple coordination sites | Higher-order structures with enhanced stability |
| Tetrazines | 4 | 1,2,4,5-tetrazine | Highly electron-deficient | Specialized magnetic and optical properties |
The synthesis of inorganic supramolecular complexes employs strategic approaches that leverage molecular self-assembly principles. Two established methodsâliquid diffusion synthesis and mechanochemical synthesisâhave proven effective for constructing these sophisticated architectures [60]. In liquid diffusion synthesis, an aqueous solution of the metal salt is layered with an ethanol solution of the organic ligand, followed by further layering with acetone and left standing at room temperature. This method typically yields X-ray quality crystals over 7-15 days, as demonstrated in the preparation of CdII halide complexes with pyrazinone, pyrimidinone, and quinazolinone ligands [60]. Alternatively, mechanochemical synthesis involves grinding cadmium(II) salts and ligands in an equimolar ratio with minimal ethanol assistance (approximately 100 µl) in a stainless steel jar with grinding balls, using a grinder operating at 25 Hz frequency [60]. This solvent-minimized approach offers environmental benefits and frequently produces identical structural outcomes to solution-based methods.
Recent advances in organic-inorganic hybrid supramolecules highlight the importance of template-directed synthesis [34]. For instance, the synthesis of a chain-like organic cation structure directing agent (L·Clâ) begins with reacting triethylenediamine with 1,2-bis(2-chloroethoxy)ethane [34]. This template cation can then be combined with various inorganic metal salts through room temperature volatilization methods to produce diverse organic-inorganic hybrid supramolecules, including {[L][HgIâ]} (1), {[L][CdIâ]} (2), {[(L)(CuâIâ)]·[CuIâ]CHâCN}â (3), {[L][CoClâ]â} (4), and {[L][Ce(NOâ)â ·(HâO)â]} (5) [34]. The circular structure of 1,4-diazabicyclo[2.2.2]octane (DABCO) provides particularly favorable geometry, with its two nitrogen atoms positioned for optimal coordination with metals, generating stable molecular complexes [34].
Comprehensive characterization of supramolecular complexes requires multiple analytical techniques to elucidate both structure and function. Single-crystal X-ray diffraction serves as the primary method for determining precise molecular geometry and supramolecular organization [60] [34]. For example, structural analyses reveal that compounds {[L][HgIâ]} (1) and {[L][CdIâ]} (2) both crystallize in the monoclinic system with the P2â/n space group, with metal ions adopting distorted tetrahedral configurations coordinated to four iodine atoms [34]. In contrast, {[(L)(CuâIâ)]·[CuIâ]CHâCN}â (3) belongs to the triclinic crystal system with the P-1 space group, forming more complex chain structures where nitrogen atoms from ligands connect to [CuIâ]â» and [CuI] ions [34].
Supplementary characterization methods include infrared spectroscopy for identifying functional groups and coordination modes, elemental analysis for verifying composition, and thermogravimetric analysis for assessing thermal stability [34]. For investigating magnetic properties, SQUID (Superconducting Quantum Interference Device) magnetometry provides extremely sensitive measurements of magnetic susceptibility and effective magnetic moment across temperature ranges from liquid helium to room temperature [59]. These instruments detect faint magnetic signals through Josephson junctions in superconducting materials, functioning as magnetic flux-to-voltage converters that enable detailed understanding of magnetic interactions in supramolecular clusters [59].
Multimodal imaging integrates two or more complementary techniques to overcome the limitations of individual modalities, combining anatomical information with superior spatial resolution and biological information at the molecular level with high sensitivity [61]. Each modality offers distinct advantages: computed tomography (CT) provides high spatial resolution, magnetic resonance imaging (MRI) offers excellent soft tissue contrast, while positron emission tomography (PET), single-photon emission computed tomography (SPECT), and optical imaging visualize biological functions at the molecular level [61]. The development of PET/CT in 1998 represented the first clinically applicable multimodal instrument, sequentially combining X-ray and radioisotope probes to generate superimposed images that fuse anatomical and functional information [61].
A significant challenge in multimodal probe design involves the differing concentration requirements for various contrast agents. For instance, CT contrast agents typically require concentrations nearly three orders of magnitude higher than those needed for PET imaging [61]. This discrepancy makes integrating all functionalities into a single molecule impractical. Furthermore, signal intensity for T1-weighted MR contrast agents does not increase linearly with Gd(III) concentration, eventually plateauing and decreasing at very high concentrations [61]. These technical constraints highlight the importance of nanoplatforms that can incorporate multiple contrast mechanisms at appropriate densities while maintaining biocompatibility and targeting specificity.
Surface engineering of nanoprobes has revolutionized bioimaging and theranostics by addressing limitations of conventional agents like photobleaching, poor tissue penetration, and nonspecific distribution [62]. Major surface engineering strategies include ligand-mediated targeting, environmental responsiveness, charge engineering, surface coating, and core-shell structure design [62]. These approaches enable high-resolution, dynamic, real-time multimodal imaging by controlling interactions at the bio-nano interface.
Optimizing surface characteristics requires careful balancing of multiple factors. Hydrophilic surface modifications increase blood half-life for extended circulation time but may negatively impact cellular uptake and targeting efficiency [61]. Similarly, surface coating materials with larger hydrodynamic volume effectively camouflage particulate imaging agents during circulation but may physically hide conjugated targeting moieties, limiting contact with receptors on cell surfaces [61]. The strategic distribution of targeting moieties on nanostructures, combined with appropriate conjugation methods, is therefore essential for maintaining imaging resolution and diagnostic quality.
Table 2: Multimodal Imaging Modalities and Their Characteristics
| Imaging Modality | Spatial Resolution | Depth Penetration | Key Contrast Mechanisms | Primary Applications |
|---|---|---|---|---|
| Magnetic Resonance Imaging (MRI) | 10-100 μm | Unlimited | Gd(III) complexes, iron oxide nanoparticles | Soft tissue contrast, anatomical imaging |
| X-ray Computed Tomography (CT) | 50-200 μm | Unlimited | Iodinated compounds, heavy metal nanoparticles | High-resolution anatomical imaging |
| Positron Emission Tomography (PET) | 1-2 mm | Unlimited | Radioisotopes (¹â¸F, â¶â´Cu, â¸â¹Zr) | Metabolic activity, molecular imaging |
| Single Photon Emission Computed Tomography (SPECT) | 1-2 mm | Unlimited | Radioisotopes (â¹â¹mTc, ¹¹¹In) | Physiological processes, molecular imaging |
| Optical Imaging | 1-10 mm | <1 cm | Fluorescent dyes, quantum dots | Superficial molecular imaging, intraoperative guidance |
| Photoacoustic Imaging | 10-500 μm | 1-5 cm | Organic dyes, gold nanoparticles | Vascular imaging, tissue oxygenation |
Active targeting strategies significantly enhance the precision of theranostic platforms by incorporating ligands that specifically bind to receptors overexpressed on target cells. This approach increases accumulation at disease sites, improves delivery efficiency, minimizes side effects, and reduces nonspecific binding [61]. Major categories of targeting moieties include amino acids (proteins, fragment domains, peptides), nucleic acids (aptamers), carbohydrates, lipids, vitamins, and polymers (hyaluronic acid) [61]. These targeting elements are conjugated to delivery systems using various methods, including direct conjugation, linker chemistry, click chemistry, and physical interactions such as ionic, hydrophobic, and host-guest interactions [61].
Each targeting moiety offers distinct advantages and limitations. Antibodies provide high selectivity and strong affinity through two binding sites but present challenges including high molar mass, expensive manufacturing, and potential immunogenicity [61]. Peptides offer lower immunogenicity, smaller size for efficient conjugation, and reduced manufacturing costs, with the arginine-glycine-aspartic acid (RGD) sequence and its cyclic forms being particularly valuable for targeting αvβ3 integrin overexpressed during tumor angiogenesis [61]. However, peptides typically exhibit relatively low binding affinity and limited stability in the bloodstream. Aptamers (15-40 base nucleic acids) provide binding affinity through three-dimensional configurations identified via Systematic Evolution of Ligands by Exponential Enrichment (SELEX), sharing similar advantages and disadvantages as peptides [61]. Low molecular weight compounds like folic acid offer cost-effective production and reproducible homogeneity, with high binding affinity for folate receptors overexpressed on tumor cells [61].
Passive targeting represents an alternative approach that exploits anatomical and physiological abnormalities in diseased tissues, particularly in tumors with defective, leaky vasculature and inadequate lymphatic drainage [61]. This phenomenon, known as the enhanced permeability and retention (EPR) effect, allows optimally sized nanostructured delivery carriers (typically 10-200 nm) to escape through pores in leaky blood vessels and accumulate selectively in target tissues [61]. The EPR effect provides extended retention time at tumor sites without additional energy consumption, forming the basis for many nanotherapeutic delivery systems.
However, passive targeting approaches face several limitations due to the diverse nature of blood vessel porosity, insufficient diffusion processes, and inhomogeneous vessel permeability even within single tumors [61]. These challenges have motivated the development of active binding systems that can specifically interact with target cells after extravasation, overcoming the constraints of passive accumulation alone.
Recent research demonstrates the practical implementation of supramolecular theranostic platforms through the development of organic-inorganic hybrid compounds with photocatalytic capabilities for environmental applications [34]. In one study, five organic-inorganic hybrid supramolecules were synthesized using a chain-like organic cation structure directing agent (L·Clâ) reacted with various inorganic metal salts through room temperature volatilization methods [34]. Compound {[(L)(CuâIâ)]·[CuIâ]CHâCN}â (3) exhibited particularly promising photocatalytic performance, achieving 92.22% degradation efficiency of tetracycline antibiotics in solution at pH = 7 with a 10 mg catalyst dose [34].
This system demonstrates the functional potential of supramolecular design, maintaining photocatalytic degradation efficiency above 86% through four consecutive cycles, indicating excellent recyclability and stability [34]. The exploration of optimal conditions for photocatalytic degradation, including solution pH, catalyst dosage, and solution temperature, provides a methodological framework for optimizing supramolecular systems for therapeutic applications. While this example addresses environmental remediation, the principles directly translate to biomedical applications where catalytic activity or responsive behavior is required within biological systems.
Table 3: Essential Research Reagents for Supramolecular Theranostic Development
| Reagent Category | Specific Examples | Function in Research | Key Characteristics |
|---|---|---|---|
| Azabenzene Ligands | Pyridine, pyrazine, pyrimidine, 1,3,5-triazine | Metal coordination building blocks | Electron-deficient, strong Ï acceptors, form stable complexes |
| Metal Salts | CdII halides, HgIâ, CuI, CoClâ, Ce(NOâ)â | Inorganic cluster formation | Variable coordination geometry, diverse magnetic/optical properties |
| Structure Directing Agents | DABCO derivatives, L·Clâ template | Control supramolecular assembly | Pre-organized geometry, multiple coordination sites |
| Targeting Moieties | RGD peptides, folate, aptamers, antibodies | Specific cellular targeting | High affinity for receptors overexpressed on target cells |
| Surface Engineering Agents | PEG, hyaluronic acid, charged polymers | Modify bioavailability and circulation | Improve solubility, reduce immunogenicity, enhance stability |
| Characterization Standards | SQUID calibration standards, XRD reference materials | Validate analytical measurements | Ensure measurement accuracy and cross-study comparability |
| Bicyclo(3.2.0)hept-1(5)-ene | Bicyclo(3.2.0)hept-1(5)-ene, CAS:10563-10-7, MF:C7H10, MW:94.15 g/mol | Chemical Reagent | Bench Chemicals |
| 1-Benzhydryl-3-nitrobenzene | 1-Benzhydryl-3-nitrobenzene | Bench Chemicals |
The clinical translation of supramolecular theranostic platforms faces several significant challenges that represent active areas of research. Biocompatibility and toxicity concerns remain paramount, particularly for inorganic clusters containing heavy metals, necessitating extensive toxicological profiling and surface modification strategies to enhance safety profiles [62] [61]. Manufacturing reproducibility at scale presents another hurdle, as the self-assembly processes crucial to supramolecular synthesis must be precisely controlled to ensure batch-to-batch consistency in complex multifunctional agents [62]. Regulatory pathways for combination products that integrate diagnostic and therapeutic functions require further clarification, as these platforms don't fit neatly into traditional medical product categories [62] [61].
Future advancement will likely focus on developing increasingly adaptive and intelligent systems that respond to multiple biological stimuli, providing dynamically controlled therapeutic action and imaging feedback [62]. The integration of artificial intelligence in molecular design and optimization promises to accelerate the discovery of novel supramolecular architectures with enhanced targeting specificity and functional integration [58]. Additionally, combination with emerging modalities like gene editing and immunotherapy may yield theranostic platforms that not only diagnose and treat but also modify disease processes at their fundamental molecular origins [61]. As these technologies mature, supramolecular chemistry will continue to provide the foundational principles for constructing increasingly sophisticated theranostic agents that blur the distinction between material and medicine.
Crystallization is a fundamental process in the synthesis of advanced materials, determining their structural integrity, phase purity, and ultimate functional properties. For researchers working with inorganic cluster supramolecular compounds, achieving precise control over crystallization parameters is not merely a technical consideration but a central requirement for obtaining materials with predictable characteristics. The process represents a delicate balance of thermodynamic and kinetic factors, where subtle alterations in reaction conditions can redirect assembly pathways toward divergent structural outcomes.
This guide examines the three most influential parameters in crystallization controlâtemperature, pH, and solvent selectionâwith specific application to inorganic cluster supramolecular compounds. These materials, which include metal-organic frameworks (MOFs), polyoxometalates, and other coordination polymers, possess complex architectures that arise from the directed assembly of metal clusters with organic linkers through supramolecular interactions [63] [64]. By systematically examining how these parameters influence crystallization pathways, researchers can transcend empirical approaches and advance toward predictive synthesis of materials with tailored properties.
Temperature serves as a primary governor of both thermodynamic stability and kinetic progression in crystallization processes. It directly influences reaction rates, solubility equilibria, and the stability of intermediate phases that often dictate final structural outcomes.
Supramolecular isomerism occurs when identical chemical compositions yield different network structures through variations in assembly pathways. Temperature manipulation provides a powerful method for accessing these structural variations. In one documented case, the same precursor (HâPTTA) produced two distinct supramolecular isomers depending solely on temperature [65]. At ambient temperature, the reaction yielded 1-L, an achiral layer structure stabilized by hydrogen bonding. When the identical reaction mixture was heated to 160°C, a different isomer (1-H) formed, consisting of chiral tubes packed into a supramolecular framework [65].
The structural differences between these isomers have significant functional implications. The chiral tubular structure of 1-H possesses internal channels capable of guest molecule inclusion, while the layered structure of 1-L does not. This demonstrates how temperature control can selectively access structural features that determine material functionality.
In the crystallization of metal-organic frameworks, temperature directly controls structural dimensionality. Research on cobalt-succinate systems revealed that progressively higher temperatures produce increasingly complex architectures [63]. At temperatures below 100°C, one-dimensional (1D) chain structures predominated. Raising the temperature to 150°C promoted two-dimensional (2D) sheet formation, while temperatures at 190°C and above yielded three-dimensional (3D) network structures [63].
This temperature-dependent dimensionality correlates with decreased water content in the final structures at higher temperatures, altering the coordination geometry around metal centers and enabling more extensive cross-linking between building units.
Temperature profoundly affects crystallization kinetics through its relationship with activation energy. The Avrami equation models this relationship, describing how the crystalline fraction (X) develops over time (t): X = 1 - e^(-ktâ¿), where k is the temperature-dependent rate constant and n is the Avrami exponent related to nucleation and growth mechanisms [66].
For viologen-based gelators, crystallization follows interface-controlled growth (nG = 1) with pre-existing nuclei (nN = 0) [66]. Lower quenching temperatures produce denser, smaller spherulitic structures due to increased supersaturation, though interestingly, the elastic modulus of the resulting gels remains largely unaffected by these morphological changes [66].
Table 1: Temperature-Induced Structural Variations in Supramolecular Compounds
| Compound System | Low Temperature Phase | High Temperature Phase | Structural Consequences |
|---|---|---|---|
| HâPTTA [65] | Achiral layer structure (1-L) at 25°C | Chiral tubular structure (1-H) at 160°C | Transition from non-porous to potentially porous architecture |
| Co-succinate [63] | 1D chains (<100°C) | 2D sheets (150°C), 3D networks (190°C) | Increased structural dimensionality and complexity |
| Viologen gels [66] | Dense, small spherulites at low T | Larger spherulites at higher T | Morphological control without altering gel elasticity |
| Cadmium nitrate + HâBDC [63] | 2D network at 95°C | 3D network at 115°C | Different supramolecular isomers based on temperature |
The pH of a reaction mixture exerts crucial influence over crystallization by modulating the protonation states of molecular building blocks. This control is particularly critical for inorganic cluster systems where ligand charge balance determines coordination possibilities with metal centers.
In organic-inorganic hybrid compounds based on vanadate chains and zinc ions, systematic pH variation produces fundamentally different structures [67]. At specific pH values, the assembly process yields a compound with helical vanadate chains connected by zinc dimers, forming a 2D chiral layer that further assembles into a 3D chiral framework. At different pH conditions, the same components form compounds containing different vanadate chain structuresâeither [(VâOââ)â´â»]â chains with non-centrosymmetric frameworks or [(VâOâ)²â»]â chains with centrosymmetric networks [67].
These structural differences originate from pH-dependent changes in vanadate speciation and the protonation state of the coordinating ligands, which alter their metal-binding capabilities. The result is supramolecular isomersâcompounds with identical chemical composition but distinct network structuresâdirected solely by pH adjustment [67].
Beyond directing overall structure, pH manipulation can control nucleation processes in colloidal crystallization systems. In the formation of methylammonium lead iodide (MAPI) perovskite thin films, precursor solution pH affects the stability of intermediate colloidal species and their subsequent nucleation behavior [68]. The transformation pathway proceeds through a four-step process: initial nanocrystal nucleation, conversion to a solvent-complex intermediate, thermal decomposition of this complex, and final recrystallization into the desired perovskite structure [68]. Each stage exhibits pH sensitivity, offering multiple control points for directing the crystallization pathway.
Table 2: pH-Dependent Structural Variations in Hybrid Compounds
| Compound System | Acidic pH Structures | Basic pH Structures | Structural Influence |
|---|---|---|---|
| Zn-Vanadate hybrids [67] | 3D chiral framework with helical chains | Non-centrosymmetric or centrosymmetric frameworks | Altered chain connectivity and crystal symmetry |
| Phosphomolybdate clusters [64] | Hydrogen-bonded frameworks | Varied nuclearity clusters | Modified supramolecular interactions and void formation |
| MAPI perovskites [68] | Affects colloidal precursor stability | Influences intermediate phase formation | Controls nucleation kinetics and film morphology |
The solvent system provides the physical medium for molecular assembly and directly participates in supramolecular interactions through solvation effects, polarity, and coordination ability.
In perovskite crystallization, solvent molecules directly incorporate into intermediate phases that template final crystal structures. During methylammonium lead iodide formation, dimethyl sulfoxide (DMSO) molecules coordinate with lead ions to create a crystalline solvent-complex intermediate (PbâIâ·2DMSO·2CHâNHâ) [68]. This intermediate only converts to the final perovskite structure upon thermal evaporation of the coordinated solvent molecules. The strength of solvent coordination directly impacts intermediate stability and conversion kinetics, with strongly coordinating solvents like DMSO producing more defined intermediate phases than weakly coordinating solvents [69] [68].
Solvent polarity determines solvation efficiency for precursor species and influences the kinetics of self-assembly processes. In metal-organic framework synthesis, common solvents like dimethylformamide (DMF), N-methyl-2-pyrrolidone (NMP), dimethyl sulfoxide (DMSO), and acetonitrile (ACN) exhibit different coordination strengths toward metal ions [63] [69]. Strongly coordinating solvents can compete with organic linkers for metal coordination sites, potentially slowing crystallization or creating alternative pathways. The stability of metal-solvent complexes directly impacts nucleation barriers and crystal growth rates [69].
Binary and ternary solvent mixtures provide enhanced control over crystallization by balancing multiple solvent properties simultaneously. These mixed systems allow researchers to fine-tune solution polarity, viscosity, and coordination ability to optimize precursor solubility and nucleation rates [69]. In perovskite systems, solvent engineering using mixed solvents has enabled the production of high-quality films with uniform morphology and reduced defect densityâkey requirements for optoelectronic applications [69].
This protocol enables systematic exploration of temperature effects on crystallization outcomes:
Reaction Mixture Preparation: Dissolve metal salt and organic ligand in appropriate solvent (e.g., DMF, water, or mixtures) to achieve typical concentrations of 1-10 mM for metal centers [63].
Vessel Selection: Distribute equal volumes of reaction solution into multiple Teflon-lined stainless steel autoclaves or sealed glass vessels suitable for the target temperature range [63] [65].
Temperature Programming: Place vessels in ovens or heating blocks set at different temperatures spanning a relevant range (e.g., room temperature, 80°C, 120°C, 160°C) [63] [65].
Reaction Time: Maintain temperatures for 24-72 hours to ensure complete reaction and crystallization.
Cooling Protocol: Implement controlled cooling rates (e.g., 5°C/h) to preserve crystal quality, particularly for high-temperature phases [65].
Product Isolation: Collect crystals by filtration, wash with mother liquor or compatible solvent, and characterize by single-crystal X-ray diffraction, powder XRD, and thermal analysis [64] [65].
This protocol describes pH-controlled crystallization for inorganic cluster compounds:
Precursor Solution: Prepare solution containing metal salt and organic ligand in water or water-solvent mixtures [67] [64].
pH Adjustment: Slowly add dilute acid (HCl, HNOâ) or base (NaOH, NHâ) to the reaction mixture while stirring continuously. Use a calibrated pH meter to monitor the process.
pH Stabilization: Allow the solution to stabilize after each pH adjustment to ensure accurate measurement before proceeding.
Reaction Execution: Transfer the pH-adjusted solution to crystallization vessels (vials for room temperature, autoclaves for hydrothermal conditions) [67].
Incubation: Maintain at constant temperature for sufficient time (typically 2-7 days) for crystal growth.
Analysis: Characterize crystalline products and compare structures across different pH values to identify structural transitions [67].
Table 3: Key Reagents for Controlled Crystallization of Supramolecular Compounds
| Reagent Category | Specific Examples | Function in Crystallization |
|---|---|---|
| Metal Salts | Cadmium nitrate, Zinc nitrate, Cobalt chloride | Provide metal centers for cluster formation and framework construction |
| Organic Linkers | Benzenedicarboxylic acid (HâBDC), 1,1â²-(1,4-butanediyl)bis(imidazole) (bbi), tetra-carboxylic acids | Bridge metal centers to form extended frameworks; functionality dictates geometry |
| Solvents | DMF, DMSO, DEF, acetonitrile, water, alcohols | Dissolve precursors; mediate assembly through polarity and coordination ability |
| pH Modifiers | HCl, HNOâ, NaOH, NHâ, organic buffers | Control protonation states of ligands and metal hydrolysis equilibria |
| Structure-Directing Agents | Alkyl amines, ionic liquids, surfactants | Template specific structures through non-covalent interactions |
| Neodymium--palladium (1/3) | Neodymium--palladium (1/3), CAS:12164-70-4, MF:NdPd3, MW:463.5 g/mol | Chemical Reagent |
Successful crystallization requires simultaneous optimization of multiple parameters rather than sequential adjustment. The following diagram illustrates the interconnected decision process for controlling crystallization outcomes:
This framework visualizes how temperature governs thermodynamic and kinetic aspects, pH directs molecular speciation and protonation states, and solvent systems template assembly through coordination and solvation effects. These parameters converge to define intermediate phases that ultimately evolve into the final crystalline product.
The controlled crystallization of inorganic cluster supramolecular compounds demands meticulous attention to the interconnected parameters of temperature, pH, and solvent environment. Through systematic investigation of these factors and their synergistic effects, researchers can advance from empirical synthesis to predictive design of functional materials. The protocols and frameworks presented here provide a foundation for developing tailored crystallization strategies that address specific structural targets and functional requirements in supramolecular chemistry and materials science.
As crystallization science continues to evolve, emerging techniques including high-throughput screening, in situ monitoring, and computational modeling will further enhance our ability to precisely control molecular assembly processes. By integrating these advanced approaches with fundamental understanding of temperature, pH, and solvent effects, researchers can unlock new structural possibilities in the design of functional supramolecular materials.
The integration of inorganic cluster supramolecular compounds into biological environments represents a frontier in materials science and drug development. These compounds, which form extended architectures through non-covalent interactions, exhibit dynamic behaviors that must be carefully managed to ensure structural integrity in complex biological media. Structural integrity is paramount for maintaining the functional properties of these materialsâsuch as controlled drug release, catalytic activity, and molecular recognitionâthroughout their therapeutic application. This technical guide provides a comprehensive framework for enhancing the stability of these sophisticated systems, with a focus on supramolecular strategies that leverage synergistic interactions, hierarchical design, and bio-inspired approaches.
The fundamental challenge lies in the inherent dynamic nature of supramolecular systems, whose integrity depends on reversible bonds rather than covalent linkages. When introduced into biological environments, these structures face additional threats from competing interactions, pH variations, enzymatic activity, and osmotic pressures. By applying principles from crystal engineering and supramolecular chemistry, researchers can design clusters with built-in resilience mechanisms, transforming potential weaknesses into strategically managed assets for responsive behavior in physiological conditions.
The structural integrity of inorganic supramolecular clusters in biological media is governed by a hierarchy of non-covalent interactions that can be engineered to provide stability under physiological conditions.
Hydrogen Bonding: This directional interaction frequently serves as the primary organizational force in supramolecular networks. In aqueous environments, the competition from water molecules necessitates the design of highly cooperative hydrogen-bonding arrays. For instance, in rhenium sulfide clusters, specific hydration patterns form robust H3O2âbridges that significantly enhance cohesion within the 3D framework, even in aqueous systems [70].
Ï-Ï Stacking: Interactions between aromatic systems provide valuable stabilization energy and can direct long-range organization. The incorporation of aromatic moieties, such as pyrazine in rhenium clusters or benzoyl groups in ornithine-derived ligands, facilitates stacking interactions that help maintain structural integrity against disruptive forces in biological media [70] [71].
Metal Coordination: The integration of metal centers, whether as part of the cluster core or as secondary linkers, provides additional stability through directional bonding with well-defined geometries. The study of GeâSââ clusters demonstrates how metal coordination can work synergistically with hydrogen bonding and Ï-Ï interactions to create robust hybrid networks [31].
Electrostatic and Van der Waals Interactions: These non-directional forces often provide the "background" stability that supplements primary directional interactions, contributing to overall packing efficiency and density.
Table 1: Hierarchy of Supramolecular Interactions Relevant to Biological Media
| Interaction Type | Energy Range (kJ/mol) | Directionality | Role in Structural Integrity | Sensitivity to Aqueous Environment |
|---|---|---|---|---|
| Hydrogen Bonding | 10-40 (moderate) | High | Primary framework formation | High (requires careful design) |
| Ï-Ï Stacking | 5-50 (moderate) | Moderate | Stabilization & organization | Low to moderate |
| Metal Coordination | 50-200 (strong) | High | Enhanced stability & direction | Low (if kinetically inert) |
| Electrostatic | 10-20 (weak) | Low | Background cohesion | High (screened by ions) |
| Van der Waals | <5 (very weak) | Low | Packing efficiency | Low |
The supramolecular synthon approach, introduced by Desiraju, provides a methodological framework for designing robust architectures [71]. Synthons are defined as structural units formed by reproducible patterns of supramolecular interactions. By identifying and utilizing reliable synthons, researchers can predictably engineer cluster-based systems with enhanced integrity.
In practice, this involves:
For ornithine-derived ligands, systematic analysis of interaction patterns has enabled the creation of a synthon library that informs the design of compounds with optimized stability profiles for biological applications [71].
Protocol 1: Hydrothermal Synthesis of Germanium Sulfide Cluster Networks [31]
Objective: To synthesize an inorganic-organic hybrid supramolecular network containing GeâSââ clusters with enhanced stability through multiple interaction types.
Materials:
Procedure:
Protocol 2: Aqueous Phase Synthesis of Rhenium Sulfide Clusters with Pyrazine [70]
Objective: To prepare supramolecular frameworks based on [{ReâSáµ¢â}(OH)ââ]â´â» cluster units with controlled hydration states for enhanced biological compatibility.
Materials:
Procedure:
Table 2: Analytical Methods for Evaluating Structural Integrity
| Technique | Information Obtained | Experimental Parameters | Relevance to Biological Media |
|---|---|---|---|
| Single Crystal X-ray Diffraction | Precise molecular geometry, interaction metrics | Temperature: 100-298 K; Resolution: <0.8 Ã | Reveals hydration patterns & interaction networks comparable to aqueous environments |
| Hirshfeld Surface Analysis | Quantification of interaction types and contributions | Based on crystallographic data; visualization of molecular surfaces | Maps interaction preferences critical for stability in competitive environments [71] |
| Circular Dichroism (CD) Spectroscopy | Monitoring supramolecular polymerization & chiral organization | Concentration: 10-500 μM; Temperature control: ±0.1°C | Tracks assembly/disassembly in near-physiological conditions [72] |
| Energy Dispersive X-ray Spectroscopy (EDS) | Elemental composition and stoichiometry | Accelerating voltage: 15-20 kV; Heavy element analysis | Confirms cluster integrity and composition after biological exposure |
| Computational Analysis (DFT) | Interaction energies, conformational preferences | B3-LYP-D3 level; solvation models | Predicts stability of synthons and their behavior in aqueous media [71] |
The management of water interactions represents perhaps the most critical factor in maintaining structural integrity in biological environments. Controlled hydration can be transformed from a destabilizing threat into a reinforcing element:
Structured Water Bridges: As demonstrated in rhenium cluster compounds, the formation of specific HâOââ» bridges between cluster units creates additional connection points that enhance framework cohesion. These bridges represent a higher-level synthon that persists in aqueous environments [70].
Hydration-Dependent Topology: The structure of supramolecular frameworks can be designed to adapt to different hydration states while maintaining integrity. Compounds 1, 2, 4, and 5 in rhenium cluster systems all maintain their structural integrity through different hydration levels (8-8.5 HâO molecules per formula unit), with the hydration shell becoming an integral part of the architecture rather than a disruptive element [70].
Synthon Conservation Across Hydration States: Effective design ensures that core synthons remain intact even as water content fluctuates. This requires the identification and utilization of hydration-resistant interaction patterns during the initial design phase.
Relying on a single interaction type creates vulnerability in biological environments where competitive binding is constant. Successful systems employ cooperative interaction networks:
Synergistic Stabilization: The GeâSââ cluster system demonstrates how hydrogen bonding (O-H···O, N-H···N, C-H···O, N-H···O, O-H···S) and Ï-Ï stacking can work cooperatively to create a robust framework where the failure of one interaction type doesn't compromise overall integrity [31].
Hierarchical Assembly: Systems can be designed where strong, directional interactions (metal coordination, primary hydrogen bonds) create a stable core framework, while weaker, more dynamic interactions (Ï-stacking, van der Waals) provide adaptability and self-healing capabilities in response to environmental fluctuations.
Biological systems offer sophisticated models for maintaining integrity in aqueous environments:
Sacrificial Bond Mechanics: Inspired by biological materials, supramolecular clusters can incorporate weaker interactions that break before primary structure failure, dissipating energy that would otherwise damage the core architecture.
Self-Healing Through Reversibility: The dynamic nature of supramolecular interactions, often viewed as a liability, can be harnessed as an asset through careful design. Systems that allow for temporary disassembly and reassembly in response to stress can recover integrity after perturbation.
Table 3: Essential Research Reagents for Supramolecular Integrity Studies
| Reagent Category | Specific Examples | Function in Research | Considerations for Biological Media |
|---|---|---|---|
| Cluster Precursors | Kâ[{ReâSáµ¢â}(OH)ââ]·8HâO, TMAâGeâSââ | Core building blocks for supramolecular frameworks | Aqueous solubility enables biological compatibility studies [70] [31] |
| Organic Linkers | Pyrazine, 4,4'-bipyridine, benzoyl derivatives | Provide interaction sites for assembly | Aromatic systems enable Ï-Ï stacking; nitrogen donors facilitate metal coordination [70] [71] |
| Modulation Agents | Ba(NOâ)â, KNOâ, Ga(NOâ)â·HâO, MgSOâ·7HâO | Control pH and crystallization conditions | Biological ions (Kâº, Mg²âº) enhance physiological relevance [70] |
| Stability Enhancers | Histidine, ornithine derivatives | Introduce multiple functional groups for diverse interactions | Amino acid-based compounds improve biological integration [71] |
| Analytical Standards | Hemicucurbit[n]urils, cyclohexanohemicucurbit[n]urils | Reference compounds for quantification | Enable calibration of analytical methods for complex mixtures [73] |
The strategic management of dynamic behavior in inorganic cluster supramolecular compounds requires a multifaceted approach that anticipates the challenges of biological environments. By employing hierarchical synthon design, controlling hydration patterns, and implementing multi-modal interaction networks, researchers can significantly enhance structural integrity without sacrificing the responsive nature that makes these systems valuable for therapeutic applications. The experimental methodologies and strategic frameworks presented here provide a foundation for developing next-generation supramolecular systems with the robustness required for advanced drug development and biomedical applications. As this field progresses, the integration of computational prediction with empirical validation will further refine our ability to design clusters with precisely controlled dynamic behavior in biological media.
The pursuit of phase purity in supramolecular architectures represents a fundamental challenge in advanced materials synthesis, particularly within the context of inorganic cluster supramolecular compounds. Phase purity refers to the isolation of a single, well-defined supramolecular architecture from a complex mixture of possible thermodynamic products, ensuring uniform structural, electronic, and functional properties across the entire material. For researchers and drug development professionals, achieving phase purity is not merely an academic exercise but a critical prerequisite for developing materials with reproducible biological activity, predictable drug release profiles, and consistent therapeutic performance. The inherent reversibility of non-covalent interactionsâincluding hydrogen bonding, Ï-Ï stacking, metal-coordination, and van der Waals forcesâintroduces significant challenges in controlling self-assembly pathways toward singular outcomes.
The field of supramolecular chemistry has evolved dramatically since the early observations of hydrogen-bonded liquid crystal structures in 1927 [74]. Today, sophisticated design principles enable the construction of complex architectures such as metal-organic frameworks (MOFs), supramolecular organic frameworks (SOFs), and various coordination polymers with precision approaching that of covalent synthesis [75] [76]. Within pharmaceutical applications, the structural fidelity of these materials directly influences critical performance parameters including drug loading capacity, release kinetics, and biological targeting. For instance, MOF-based drug delivery systems demonstrate loading capacities up to 25 wt%âsignificantly higher than traditional liposomal systemsâbut this advantage depends entirely on achieving consistent structural perfection across batches [76]. This technical guide comprehensively addresses the most advanced methodologies for ensuring phase purity in supramolecular architectures, with particular emphasis on techniques relevant to inorganic cluster compounds and their applications in pharmaceutical development.
The pathway toward phase-pure supramolecular materials encounters several intrinsic obstacles rooted in the nature of non-covalent synthesis. Unlike covalent bond formation, supramolecular assembly involves reversible interactions that continually form and dissociate under reaction conditions, leading to potential kinetic traps and polymorphic outcomes. For inorganic cluster supramolecular compounds, additional complexity arises from the dynamic behavior of metal-ligand coordination, which can adopt multiple coordination geometries and stoichiometries depending on reaction conditions [76].
A primary challenge lies in the energy landscape of supramolecular assembly, where multiple minima often correspond to structurally distinct yet energetically similar aggregates. The presence of these metastable states frequently results in mixtures of architectures rather than a single phase-pure material. This problem is particularly pronounced in multicomponent crystals, where variations in proton transfer states can lead to distinct crystalline formsâas demonstrated in multicomponent crystals of imidazole-based drugs where cocrystal, salt, and cocrystal of salt forms can coexist from similar starting materials [77].
Furthermore, the dynamic equilibria governing supramolecular systems render them exquisitely sensitive to subtle variations in synthesis parameters including temperature, concentration, solvent polarity, ionic strength, and even mixing rates. For pharmaceutical applications, this sensitivity poses significant scale-up challenges, as small changes in process conditions can yield different architectural polymorphs with potentially divergent biological behaviors and therapeutic efficacy.
Controlling the balance between thermodynamic and kinetic factors represents the foundational principle for achieving phase purity in supramolecular architectures. Thermodynamic control leverages reversible assembly processes to allow error correction and progression toward the most stable architectural form, typically achieved through prolonged reaction times at elevated temperatures or in self-correcting solvent systems. Conversely, kinetic control utilizes rapid assembly under conditions that favor one pathway over alternatives, effectively trapping a desired architecture before competing forms can nucleate and grow.
The hydro/solvothermal method exemplifies a thermodynamically-controlled approach widely employed for synthesizing phase-pure MOFs. This technique subjects metal ions and organic ligands to elevated temperatures and pressures in sealed vessels, enhancing solubility and reversibility to enable the formation of highly crystalline, thermodynamically stable products [76]. For instance, phase-pure UiO-66 and MIL-series MOFs with well-defined porosityâcritical for drug loading applicationsâare routinely obtained through carefully optimized solvothermal conditions [76]. Recent modifications, such as photo-induced solvothermal synthesis, have significantly reduced reaction times while maintaining phase purity, demonstrating how traditional methods can be enhanced for greater efficiency [76].
In contrast, microwave-assisted synthesis represents a kinetically-influenced approach where rapid, uniform heating promotes simultaneous nucleation and fast growth, often yielding smaller crystalline domains with reduced defects and improved phase purity compared to conventional heating [76]. The preparation of MIL-53(Fe) via microwave irradiation illustrates how this technique can produce phase-pure materials with consistent porosity in significantly reduced timeframes [76].
Recent technological advances have introduced sophisticated strategies for achieving unprecedented levels of phase purity in complex supramolecular architectures:
Template-Directed Synthesis utilizes molecular or supramolecular templates that preorganize components into specific geometries, guiding assembly toward a single architectural outcome. This approach mirrors biological systems where templates direct the formation of complex structures with high fidelity.
Seed-Mediated Crystallization introduces pre-formed crystalline nuclei of the target architecture into supersaturated solutions, bypassing spontaneous nucleation events that often yield polymorphic mixtures. This technique is particularly valuable for scaling up phase-pure materials while maintaining structural consistency.
Automated High-Throughput Screening employs robotics and parallel synthesis to rapidly explore vast parameter spaces (solvent composition, concentration, temperature, stoichiometry), identifying conditions that selectively yield phase-pure products [78]. This data-driven approach has dramatically accelerated the discovery of optimal synthesis conditions for complex supramolecular systems.
The emerging application of artificial intelligence in predicting synthesis outcomes represents a paradigm shift in phase purity achievement. Machine learning algorithms trained on structural and synthetic data can now predict optimal building blocks and conditions for targeting specific supramolecular architectures with high purity [76].
Table 1: Synthesis Techniques for Achieving Phase Purity in Supramolecular Architectures
| Technique | Control Mechanism | Key Parameters | Representative Architectures | Phase Purity Advantages |
|---|---|---|---|---|
| Hydro/Solvothermal | Thermodynamic | Temperature, pressure, time | UiO-66, MIL-100, ZIF-8 | High crystallinity, defect correction |
| Microwave-Assisted | Kinetic/Thermodynamic | Power, irradiation time, temperature | MIL-53(Fe), HKUST-1 | Uniform nucleation, reduced polymorphism |
| Slow Evaporation | Thermodynamic | Solvent composition, temperature, gradient | Multicomponent pharmaceutical crystals [77] | Selective nucleation, gradual assembly |
| Diffusion Methods | Kinetic | Concentration gradient, interface control | Supramolecular organic frameworks [75] | Spatial separation of nucleation and growth |
| Electrochemical | Kinetic | Current density, potential, electrolyte | Conductive MOFs, metallosupramolecules | Controlled deposition, oriented growth |
Rigorous characterization is indispensable for verifying phase purity in supramolecular architectures, requiring a multi-technique approach to assess structural homogeneity at multiple length scales. Single-crystal X-ray diffraction remains the gold standard for unambiguous determination of molecular packing and long-range order, providing definitive evidence of phase purity when sufficient crystal quality is achieved [77]. For pharmaceutical cocrystals involving imidazole-based drugs, this technique has revealed subtle differences in proton transfer states and hydrogen bonding patterns that distinguish phase-pure forms from polymorphic mixtures [77].
Powder X-ray diffraction (PXRD) serves as a workhorse technique for bulk phase assessment, where the experimental pattern is compared against simulated patterns from single-crystal data. Divergences between observed and calculated patterns indicate the presence of impurities or polymorphic contamination. Hirshfeld surface analysis provides visual assessment of intermolecular interactions and their consistency throughout the crystal lattice, offering insights into the uniformity of supramolecular synthons [75] [77]. This approach has proven particularly valuable for characterizing phase-pure supramolecular organic frameworks based on chiral natural compounds, where complex interaction patterns must remain consistent throughout the architecture [75].
Thermal methods including thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) provide complementary evidence of phase purity through sharp, single transitions corresponding to architectural decomposition or phase changes. Broader or multiple transitions suggest the presence of structural variants or impurities. For supramolecular materials intended for pharmaceutical applications, spectroscopic techniques including solid-state NMR, IR, and Raman spectroscopy offer additional verification of chemical environment consistency throughout the material.
Table 2: Characterization Techniques for Verifying Phase Purity
| Technique | Structural Information | Detection Limits for Impurities | Sample Requirements | Applications in Supramolecular Chemistry |
|---|---|---|---|---|
| Single-Crystal XRD | Atomic coordinates, packing | ~5% (with careful refinement) | Single crystals >50μm | Definitive structure determination [77] |
| Powder XRD | Bulk crystalline phase | 1-5% | Polycrystalline powder | Phase identification, quantification |
| Solid-State NMR | Molecular environment, dynamics | 1-5% | ~50-100mg | H-bonding verification, guest incorporation |
| TGA/DSC | Thermal stability, phase transitions | 2-5% | 5-20mg | Solvent content, framework stability |
| Hirshfeld Surface Analysis | Intermolecular interactions | Qualitative | Computational (crystal structure) | Interaction fingerprinting [77] |
| Nâ Adsorption | Porosity, surface area | 5-10% | 50-200mg | BET surface area, pore volume distribution |
This protocol describes the synthesis of phase-pure MOFs via hydrothermal methods, adapted from procedures used for UiO-66 and MIL-100 synthesis [76]:
Precursor Preparation: Dissolve metal salt (e.g., ZrClâ, FeClâ·6HâO) and organic ligand (e.g., terephthalic acid, trimesic acid) in separate portions of the chosen solvent (typically DMF, water, or their mixtures). Use stoichiometric ratios determined from preliminary screening experiments.
Reaction Mixture: Combine the solutions in a Teflon-lined stainless steel autoclave, ensuring thorough mixing. For zirconium-based MOFs, typical conditions use ZrClâ:terephthalic acid:DMF molar ratios of 1:1:300.
Thermal Treatment: Seal the autoclave and heat to the target temperature (typically 100-120°C for UiO-66, 150-180°C for MIL-100) for 12-72 hours. The extended time at elevated temperature enables error correction and progression toward the thermodynamically most stable phase.
Product Recovery: Cool the autoclave slowly to room temperature (2-5°C/hour) to prevent defect formation. Collect crystalline product by filtration or centrifugation.
Purification: Wash sequentially with the reaction solvent, then with a miscible solvent (typically methanol or acetone) to remove unreacted precursors and solvent molecules trapped in pores. Exchange solvents gradually to prevent framework collapse.
Activation: Remove residual solvent under reduced pressure at elevated temperature (typically 150-200°C for 12-24 hours) to obtain the activated, porous framework.
Phase purity verification: Compare PXRD pattern with simulated pattern from reference structure; assess surface area and pore size distribution by Nâ adsorption at 77K; confirm elemental composition by ICP-OES and elemental analysis.
This protocol for obtaining phase-pure multicomponent crystals of imidazole-based drugs is adapted from procedures described for metronidazole, ketoconazole, and miconazole cocrystals [77]:
Solution Preparation: Dissolve equimolar amounts (0.05-0.5 mmol) of active pharmaceutical ingredient (e.g., metronidazole) and coformer (e.g., trithiocyanuric acid) in an appropriate solvent system. For metronidazole-TTCA cocrystals, a water-methanol (1:2, v/v) system has proven effective [77].
Filtration: Filter the solution through a 0.45μm membrane to remove particulate matter that could seed uncontrolled nucleation.
Evaporation Setup: Transfer the filtrate to a clean crystallization vessel covered with perforated parafilm to control evaporation rate. For systems prone to oiling out, consider using anti-solvent vapor diffusion instead.
Controlled Crystallization: Place the vessel in a vibration-free environment at constant temperature (typically 4°C for refrigerator evaporation or 25°C for room temperature evaporation). Maintain stable conditions for 1-2 weeks until crystals form.
Crystal Selection: Visually inspect crystals under polarized light to identify morphologically uniform specimens. For definitive characterization, select single crystals with well-defined faces and absence of inclusions.
Characterization: Determine crystal structure by single-crystal X-ray diffraction. Verify bulk phase purity by comparing experimental PXRD pattern with pattern simulated from single-crystal data.
Critical parameters for success: solvent composition, evaporation rate, temperature stability, and stoichiometric control. Even minor deviations can result in polymorphic contamination or solvate formation.
The following table details key research reagents and materials essential for achieving phase purity in supramolecular architecture synthesis, particularly focused on inorganic cluster compounds and pharmaceutical cocrystals:
Table 3: Research Reagent Solutions for Supramolecular Synthesis
| Reagent/Material | Function | Application Examples | Phase Purity Considerations |
|---|---|---|---|
| Trithiocyanuric acid (TTCA) | Hydrogen-bonding coformer | Multicomponent crystals with imidazole drugs [77] | Forms specific NâHâ¯N and NâHâ¯S interactions directing assembly |
| Metal salts (ZrClâ, FeClâ, HfClâ) | Metal cluster precursors | MOF synthesis (UiO-66, MIL-100) [76] | Purity affects nucleation kinetics; hygroscopic salts require careful handling |
| Modulated synthesis additives (Benzoic acid, acetic acid) | Coordination modulators | Defect control in UiO-66, HKUST-1 | Competes with linker binding, enhances reversibility, improves crystallinity |
| High-boiling point solvents (DMF, DEF, DMA) | Reaction medium for solvothermal synthesis | MOF crystallization [76] | Polarity and coordination ability direct structure formation |
| Deuterated solvents (DMSO-dâ, CDClâ) | NMR analysis | Solution-state characterization of intermediates | Enables monitoring of assembly processes and intermediate species |
| Silanized glassware | Non-stick surface | Slow evaporation crystallization | Prevents uncontrolled heterogeneous nucleation |
The following diagram illustrates the integrated experimental and computational workflow for achieving and verifying phase purity in supramolecular architectures:
The following diagram outlines the quality control pathway specifically tailored for pharmaceutical supramolecular materials:
The critical importance of phase purity in supramolecular architectures finds its ultimate validation in pharmaceutical applications, where structural consistency directly translates to predictable therapeutic performance. Metal-organic frameworks with precisely defined porosity enable controlled drug loading and release profiles that depend entirely on maintaining architectural integrity across batches [76]. For instance, phase-pure iron-based MIL-100 demonstrates drug loading capacities up to 25 wt%âsignificantly higher than traditional liposomal systemsâbut this performance advantage diminishes rapidly with structural defects or polymorphic contamination [76].
In multicomponent pharmaceutical crystals, phase purity ensures consistent dissolution rates and bioavailabilityâcritical parameters for regulatory approval and clinical efficacy. The research on imidazole-based drugs with trithiocyanuric acid demonstrates how different proton transfer states (cocrystal vs. salt) yield distinct solid forms with unique physicochemical properties [77]. Only through rigorous phase control can researchers ensure that the optimal form is consistently produced for formulation development.
Advanced stimuli-responsive drug delivery systems represent perhaps the most sophisticated application of phase-pure supramolecular architectures in pharmaceuticals. MOF-based systems that respond to tumor microenvironment cues (pH, glutathione concentration) or external triggers (light, magnetic fields) require exact structural control to achieve precise release kinetics [76]. The integration of artificial intelligence in MOF synthesis planning promises to accelerate the discovery of phase-pure materials optimized for specific therapeutic applications, potentially revolutionizing targeted cancer therapies and other precision medicine approaches [76].
The pursuit of phase purity in supramolecular architectures represents a central challenge in advancing the field from fundamental research to practical applications, particularly in pharmaceutical development. While significant progress has been made in understanding the principles governing supramolecular assembly, achieving consistent structural fidelity across diverse systems remains a formidable objective. The techniques outlined in this guideâfrom controlled synthesis methodologies to rigorous characterization protocolsâprovide a foundation for researchers seeking to isolate well-defined supramolecular architectures with the reproducibility required for drug development.
Future advancements will likely emerge from several promising directions. The integration of artificial intelligence and machine learning in synthesis prediction represents a paradigm shift, potentially enabling researchers to navigate complex energy landscapes and identify optimal conditions for phase-pure assembly [76]. The growing emphasis on automated high-throughput screening accelerates empirical optimization while generating valuable datasets that feed back into predictive models [78]. Additionally, advanced in situ monitoring techniques provide real-time observation of assembly processes, offering unprecedented insights into nucleation and growth mechanisms that determine architectural outcomes.
For pharmaceutical scientists, the ongoing convergence of supramolecular chemistry with materials informatics promises an era of rationally designed, phase-pure architectures with tailored therapeutic properties. As these capabilities mature, supramolecular compounds will increasingly transition from laboratory curiosities to reliable components of advanced pharmaceutical formulations, fulfilling their potential to address complex challenges in drug delivery and targeted therapy.
The synthesis of inorganic cluster supramolecular compounds represents a frontier in materials science, with applications ranging from catalysis and drug delivery to energy storage [79]. These complex architectures are constructed through the careful orchestration of non-covalent interactionsâincluding hydrogen bonding, van der Waals forces, and electrostatic attractionsâwhich govern molecular self-assembly [80] [79]. Unlike traditional covalent synthesis, supramolecular assembly is highly sensitive to reaction conditions, where subtle changes in parameters can redirect pathways toward different structural outcomes and material properties.
Achieving precise control over these systems requires a deep understanding of three fundamental reaction parameters: concentration, solvent polarity, and ionic strength. Concentration gradients directly influence nucleation kinetics and thermodynamic outcomes [80]. Solvent polarity mediates molecular recognition events and stabilization of intermediates through hydrogen-bonding networks and dielectric effects [79] [81]. Ionic strength modulates electrostatic screening and can stabilize or disrupt delicate charge balances essential for cluster integrity [79]. This technical guide provides a comprehensive framework for optimizing these parameters within the context of inorganic cluster supramolecular chemistry, supported by quantitative data, experimental protocols, and visualization tools to enable reproducible and targeted synthesis.
Supramolecular compounds are characterized by their reliance on non-covalent interactions, which are significantly weaker and more environment-dependent than covalent bonds. The directionality and strength of hydrogen bonds, Ï-Ï interactions, and metal-ligand coordination are profoundly influenced by the reaction medium [80] [79]. For instance, the extended hydrogen-bonding networks that stabilize porphyrin microsheets can be disrupted by competitive solvents, while the self-assembly of metal-organic frameworks (MOFs) depends on solvent-mediated coordination kinetics [80]. The inherent reversibility of these interactions enables self-correction and pathway selection, but also introduces sensitivity to parameter fluctuations that can lead to kinetic trapping or polymorphic outcomes.
The partition coefficient (Kpc) provides a fundamental metric for predicting molecular distribution between immiscible phases, quantitatively expressing a solute's affinity for polar versus non-polar environments [82]. Calculated as Kpc = concentration in organic phase / concentration in aqueous phase, this dimensionless parameter offers insights into molecular polarity and solvation thermodynamics. For supramolecular synthesis, Kpc values inform solvent selection to control reactant solubility, intermediate stabilization, and ultimate product distribution. Molecular structure directly determines Kpc; non-polar groups (e.g., alkyl chains) increase organic phase affinity, while polar functional groups (e.g., -OH, -COOH) enhance aqueous solubility [82]. These principles extend to supramolecular systems where compartmentalization and phase transfer can direct assembly pathways.
Table 1: Factors Influencing Partition Coefficient (Kpc) and Assembly Outcomes
| Factor | Impact on Kpc | Effect on Supramolecular Assembly |
|---|---|---|
| Solute Polarity | Higher polarity decreases Kpc | Enhances aqueous phase assembly; promotes polar interfaces |
| Solvent Pair Selection | Varies with solvent combination | Determines solute availability and concentration in each phase |
| Temperature | Temperature-dependent solubility changes | Alters equilibrium positions and thermal energy for reorganization |
| pH | Impacts ionization state of solutes | Controls protonation state and hydrogen-bonding capability |
| Ionic Strength | Affects solubility via salting-out/in | Screens electrostatic interactions between charged building blocks |
Concentration directly governs the frequency of molecular collisions and the thermodynamics of self-assembly processes. In supramolecular chemistry, high dilution conditions often favor intramolecular cyclization over polymerization by reducing intermolecular encounter rates, a principle exploited in macrocycle synthesis [80]. Conversely, higher concentrations typically promote extended assemblies and precipitation of thermodynamic products. However, traditional batch reactors often exhibit concentration gradients that lead to heterogeneous mixing and inconsistent results [80]. Localized regions of high concentration can trigger uncontrolled nucleation, yielding polydisperse products or kinetic intermediates that fail to reorganize into the desired structures.
Flow chemistry platforms address concentration gradient limitations by providing continuous, homogeneous mixing with precise residence time control [80] [83]. Microfluidic reactors with channel dimensions below 1 mm enable exact spatial and temporal control over reagent mixing, allowing supramolecular chemists to establish precise concentration profiles that direct assembly along specific pathways [80]. For instance, Numata et al. achieved the formation of uniform porphyrin microsheets under microflow conditions that were inaccessible in batch reactors, where only amorphous material formed [80]. This approach enables the targeting of non-equilibrium structures through careful manipulation of local concentration gradients and mixing efficiencies.
Table 2: Concentration Optimization Techniques in Supramolecular Synthesis
| Technique | Mechanism | Application Example | Optimal Use Cases |
|---|---|---|---|
| High-Dilution Methods | Reduces intermolecular collision frequency | Macrocycle formation [80] | Cyclic structure synthesis |
| Slow Addition Techniques | Controls introduction rate of precursors | Porphyrin assembly [80] | Hierarchical structure growth |
| Flow Chemistry | Continuous mixing with defined concentration gradients | Microsheet formation [80] | Non-equilibrium structure targeting |
| Template-Directed Assembly | Uses molecular templates to pre-organize components | Metal-organic frameworks [80] | Porous material synthesis |
Materials: Syringe pumps with precision flow control, microfluidic chip (Y-shaped mixer with reaction channel), stock solutions of building blocks, appropriate solvent system, in-line spectrometer or sampling ports for analysis.
Method:
Data Interpretation: Shorter residence times with higher concentrations often yield kinetic products, while extended residence at lower concentrations favors thermodynamic minima. The transition between regimes can be identified by sharp changes in spectroscopic signatures or morphology.
Solvent polarity directly mediates molecular recognition events through its influence on solvation shells, hydrogen-bonding networks, and dielectric environments. The principle of "like dissolves like" manifests strongly in supramolecular chemistry, where polar solvents can disrupt hydrogen-bonding motifs essential for structure direction, while non-polar solvents may insufficiently solubilize ionic precursors [82] [79]. Research has demonstrated that solvent-controlled isomerism can dictate cluster configuration in metal carbonyl systems, where specific solvents stabilize distinct isomeric forms through differential solvation of molecular domains [84]. Similarly, the use of fluorinated alcohols like hexafluoro-2-propanol (HFIP) and perfluoro-tert-butanol (PFTB) creates extended hydrogen-bonding networks that template polyene cyclization, mimicking enzymatic cavity control [81].
Ionic liquids (ILs) represent a transformative solvent platform for supramolecular synthesis due to their tunable physicochemical properties [79]. These salts with melting points below 100°C combine low volatility with extensive hydrogen-bonding capability, electrostatic interactions, and structural organization that can direct nanomaterial growth. The 1-alkyl-3-methylimidazolium [Cnmim]+ family, for example, offers multiple interaction modes: acidic protons at the C2 position for hydrogen bonding, aromatic cores for Ï-Ï interactions, and alkyl chains for van der Waals contacts [79]. This multifaceted interaction portfolio enables precise control over nucleation and growth processes, often yielding products with superior phase purity and morphological control compared to conventional solvents.
Deep Eutectic Solvents (DES) extend these advantages through eutectic mixtures of hydrogen-bond donors and acceptors (e.g., choline chloride-urea), creating designer solvents with specific polarity and coordination properties [79]. The ionic strength of DES, though lower than pure ILs due to molecular components, remains sufficient to modulate electrostatic interactions during self-assembly while providing biocompatibility and sustainability advantages.
Materials: Selected solvent series (e.g., hexane, toluene, dichloromethane, ethanol, water), ionic liquids (e.g., [C4mim][BF4], [C4mim][PF6]), pre-formed inorganic clusters, analytical equipment (UV-Vis, NMR, DLS).
Method:
Data Interpretation: Optimal solvent polarity typically manifests as maximum cluster stability (no precipitation or decomposition) with minimal polydispersity. Ionic liquids often enable synthesis at higher temperatures than molecular solvents, potentially increasing reaction rates without compromising product quality.
Ionic strength (I) quantitatively describes the concentration of ions in solution, calculated as I = ½ΣciZi², where ci is the ion concentration and Zi is the charge. In supramolecular chemistry, ionic strength controls the Debye screening length, which determines the effective range of electrostatic interactionsâa critical factor for charged building blocks and metal-ligand coordination [79]. Elevated ionic strength compresses the electrical double layer around charged species, potentially destabilizing electrostatic assemblies while sometimes enhancing hydrophobic interactions. This delicate balance requires careful optimization, as both insufficient and excessive ionic strength can prevent target structure formation.
Ionic liquids provide a unique approach to ionic strength control, as they simultaneously serve as solvent and electrolyte [79]. Their high intrinsic ionic concentration (typically 3-5 M) creates a strong electrostatic environment without requiring additional salts that might introduce competing ions or precipitation issues. The tunable nature of ILs allows researchers to design cations and anions that provide specific electrostatic environments while participating in structure direction through secondary interactions. For instance, imidazolium-based ILs can engage in Ï-stacking with aromatic building blocks while providing counterions that modulate coordination geometry around metal centers [79].
Materials: Inorganic salt solutions (NaCl, KCl, NH4OAc, etc.), buffer solutions, ionic liquids (optional), building block solutions, conductivity meter.
Method:
Data Interpretation: Plot product yield or stability against ionic strength to identify optimal ranges. Note that the effects of specific ions (Hofmeister series) may be significant, requiring comparison of different salt types at equivalent ionic strengths.
The complex interplay between concentration, solvent polarity, and ionic strength necessitates integrated optimization strategies rather than one-factor-at-a-time approaches. High-throughput screening (HTS) enables rapid exploration of this multidimensional parameter space through parallel experimentation [80] [83]. Automated liquid handling systems can prepare hundreds of microreactions with systematically varied conditions, dramatically accelerating the identification of optimal synthesis windows. For example, HTS approaches have led to the discovery of 33 new organic cages through automated workflows that would be impractical with manual methods [83].
Diagram 1: Parameter Optimization Workflow (82 characters)
Flow chemistry platforms provide exceptional control over all three key parameters simultaneously [80] [83]. Concentration gradients can be precisely engineered through controlled mixing of streams, solvent composition can be varied continuously through proportioning valves, and ionic strength can be modulated through in-line dosing of electrolyte solutions. The enhanced mixing and heat transfer in microreactors improve reproducibility while enabling access to non-equilibrium structures through precise control of assembly pathways [80]. For instance, coordination polymers and metal-organic frameworks with unusual morphologies have been synthesized under flow conditions that cannot be replicated in batch [80].
Diagram 2: Integrated Flow Reactor System (37 characters)
Table 3: Key Research Reagent Solutions for Supramolecular Synthesis
| Reagent/Material | Function | Application Notes |
|---|---|---|
| Ionic Liquids | Tunable solvent with integrated ionic strength control | [Cnmim][X] series allows systematic polarity variation [79] |
| Fluorinated Alcohols | Hydrogen-bonding templates for controlled cyclization | HFIP and PFTB create supramolecular solvent structures [81] |
| Microfluidic Reactors | Precision control of concentration and mixing | Enables kinetic pathway targeting [80] |
| High-Throughput Screening Platforms | Rapid exploration of parameter space | Automated workflows for concentration/solvent screening [80] [83] |
| Partition Coefficient Standards | Reference compounds for solvent polarity assessment | Validates solvent system selection [82] |
| Spectroscopic Tags | In-situ monitoring of assembly processes | NMR, UV-Vis probes for real-time kinetics [80] |
The targeted synthesis of inorganic cluster supramolecular compounds demands precise optimization of concentration, solvent polarity, and ionic strength parameters. These factors collectively govern the non-covalent interactions that direct molecular self-assembly along specific pathways toward functional materials. Advanced tools including ionic liquids, flow chemistry, and high-throughput screening have transformed our ability to navigate this complex parameter space efficiently. By adopting the integrated workflows and experimental protocols outlined in this technical guide, researchers can accelerate the discovery and optimization of supramolecular compounds with enhanced control over structure and properties. The continued integration of computational prediction with automated experimentation promises to further streamline this process, ultimately enabling the predictive synthesis of complex supramolecular architectures for applications across drug development, catalysis, and advanced materials.
The rational design of supramolecular frameworks based on inorganic clusters represents a frontier in materials science, with profound implications for catalysis, drug delivery, and functional materials development. Central to this design process is the mastery of two critical factors: hydration and protonation. These seemingly subtle variables exert profound influence over the final topology, stability, and properties of the resulting frameworks [70] [85]. Within the context of inorganic cluster supramolecular compound synthesis, controlling these factors enables researchers to navigate complex energy landscapes toward desired structural outcomes.
Hexarhenium chalcogenide clusters have emerged as particularly versatile building blocks due to their chemical versatility, unique redox properties, and luminescence in the red-near IR window [70]. When functionalized with organic ligands such as pyrazine, these clusters form extended supramolecular architectures whose final form is intimately tied to their aqueous synthesis environment. The hydration rate and protonation state of cluster units directly govern the supramolecular synthons that direct assemblyâthe specific patterns of hydrogen bonding, Ï-Ï stacking, and other non-covalent interactions that define the final framework topology [70] [86]. This technical guide examines the fundamental principles and experimental methodologies for controlling these critical parameters to achieve predictable framework architectures with enhanced stability.
Hydration effects operate at multiple levels in supramolecular framework formation. At the molecular level, water molecules participate directly in the hydrogen-bonding network that stabilizes the crystal structure, sometimes forming distinctive structural motifs such as HâOââ» bridges between cluster units [70]. At the supramolecular level, the overall hydration rateâthe amount of water present during assemblyâdetermines the dominance of hydrogen bonding relative to other interaction modes. Research demonstrates that H-bonding interactions become increasingly dominant as water content rises during framework formation [70].
The stability of supramolecular structures in aqueous environments further underscores hydration's importance. In G4-structure-based supramolecular hydrogels, the disorder and strength of the hydration effect proves critical for maintaining structural integrity [85]. The ability of water molecules to access functional groups involved in hydrogen bonding can lead to kinetic trapping or facilitate transformation to thermodynamic products, highlighting hydration's role in directing assembly pathways [86].
Protonation directly influences the overall charge distribution of cluster units, which in turn affects their assembly behavior. In rhenium cluster systems, varying degrees of protonation on remaining hydroxyl groups yield cluster units with different chargesâfrom neutral to anionicâthat require different counterions for charge balance [70]. This charge variation enables the incorporation of different metal cations (e.g., Kâº, Mg²âº) that further influence framework topology through coordination interactions [70].
The degree of protonation also determines the balance between different supramolecular interactions. Neutral cluster units tend to form frameworks governed primarily by hydrogen bonding, while charged units introduce Coulomb forces that work in concert with stacking interactions and hydrogen bonding [70]. This interplay between protonation state and interaction hierarchy provides a powerful lever for controlling framework properties and stability.
A comprehensive investigation into pyrazine-functionalized rhenium sulfide clusters revealed how controlled hydration and protonation yield distinct framework topologies [70]. Researchers synthesized five unique compounds from the same starting cluster salt, Kâ[{ReâSáµ¢â}(OH)ââ]·8HâO, by varying reaction conditions to manipulate protonation states and hydration levels.
The table below summarizes the five obtained compounds and their key characteristics:
Table 1: Structural Characteristics of Rhenium-Pyrazine Cluster Compounds
| Compound | Cluster Unit Formula | Configuration | Key Supramolecular Interactions | Protonation State |
|---|---|---|---|---|
| (1) | [trans-{ReâSáµ¢â}(Pz)ââ(OH)ââ(HâO)ââ] | trans | H-bonding, HâOââ» bridges | Neutral |
| (2) | [cis-{ReâSáµ¢â}(Pz)ââ(OH)ââ(HâO)ââ] | cis | H-bonding, HâOââ» bridges | Neutral |
| (3) | (NOâ)cis-{ReâSáµ¢â}(Pz)ââ(OH)â(HâO)ââ·3HâO | cis | Stacking, H-bonding only | Anionic |
| (4) | [Mg(HâO)â]â.â [cis-{ReâSáµ¢â}(Pz)ââ(OH)ââ(HâO)â]·8.5HâO | cis | H-bonding, HâOââ» bridges | Anionic |
| (5) | K[cis-{ReâSáµ¢â}(Pz)ââ(OH)ââ(HâO)â]·8HâO | cis | H-bonding, HâOââ» bridges | Anionic |
This systematic study demonstrated that the nature of synthons governing cluster unit assembly depends directly on the hydration rate of the unit [70]. Neutral building blocks (1 and 2) formed frameworks governed by hydrogen bonding, while charged units (3, 4, and 5) relied on multiple interactions including H-bonding, Ï-Ï stacking, and Coulomb forces. The presence or absence of distinctive HâOââ» bridges further highlighted how water organization contributes to structural cohesion.
The influence of hydration extends beyond initial framework assembly to include post-synthetic structural transformations. Research on bitartrate supramolecular structures on Cu(110) surfaces revealed that water initially decorates metal channels through hydrogen bonding with exposed oxygen ligands [87]. At elevated temperatures, water molecules insert into the structure, break existing intermolecular hydrogen bonds, and fundamentally change molecular adsorption sites and footprints [87].
This work demonstrated that hydration of polar metal-adsorbate ligands represents a dominant factor during surface hydration or self-assembly from solution. The transformation from a hydrophilic channel/hydrophobic row structure to a fully hydrated interface illustrates how water can progressively reshape supramolecular architectures through competitive hydrogen bonding.
The following detailed methodology enables the synthesis of hydration- and protonation-tuned rhenium cluster frameworks [70]:
Table 2: Key Research Reagents for Hydration and Protonation Control in Cluster Frameworks
| Reagent/Material | Function in Synthesis | Specific Example |
|---|---|---|
| Hexarhenium Chalcogenide Cluster Salts | Primary building blocks for framework construction | Kâ[{ReâSáµ¢â}(OH)ââ]·8HâO provides the cluster core for functionalization [70] |
| Nitrogen-Containing Heterocycles | Apical ligands for functionalization and directing assembly | Pyrazine enables apical ligand exchange and directs framework topology through stacking interactions [70] |
| Alkaline and Alkaline Earth Salts | Counterions for charged frameworks and structure direction | MgSOâ·7HâO provides Mg²⺠counterions for anionic cluster units in Compound (4) [70] |
| pH Modifiers | Control protonation states of cluster units | KOH solutions adjust overall pH, affecting hydroxyl group protonation in cluster units [70] |
| Structure-Directing Solvents | Mediate supramolecular interactions through competitive binding | Water molecules participate directly in H-bond networks and HâOââ» bridge formation [70] [85] |
The following diagram illustrates the systematic workflow for synthesizing hydration- and protonation-controlled rhenium cluster frameworks:
This diagram illustrates the relationship between hydration/protonation states and the resulting framework stability:
The controlled manipulation of hydration and protonation states represents a powerful strategy for directing the assembly and stabilizing the architecture of inorganic cluster-based supramolecular frameworks. Experimental evidence demonstrates that these factors determine the hierarchy of supramolecular interactionsâhydrogen bonding, Ï-Ï stacking, and Coulomb forcesâthat govern final framework topology [70]. The methodologies outlined herein provide researchers with precise synthetic control over these parameters, enabling the rational design of materials with tailored properties for applications in catalysis, drug delivery, and functional materials. As research advances, the fundamental principles governing hydration and protonation effects will continue to enable increasingly sophisticated control over supramolecular architecture and stability.
Inorganic Cluster Supramolecular (ICS) compounds represent a frontier in materials science, where metal clusters are organized via supramolecular interactionsâsuch as electrostatic forces, hydrogen bonding, and van der Waals forcesâinto highly ordered, functional architectures. [34] This paradigm shift moves beyond traditional covalent synthesis to leverage "chemistry beyond the molecule," enabling the design of systems with dynamic, reversible, and stimuli-responsive characteristics. [47] [34] The application of these materials in catalysis is a primary focus of contemporary research, aiming to surpass the limitations of conventional homogeneous and heterogeneous catalysts. This review provides an in-depth technical comparison of ICS-derived nanocatalysts against traditional catalytic systems, focusing on their performance in model reactions, underpinned by experimental data and mechanistic insights relevant to drug development and industrial chemistry.
ICS compounds are a class of materials where inorganic metal clusters, such as [HgI4]2-, [CdI4]2-, or [Cu2I3]-, are directionally assembled using organic cationic templates or ligands through non-covalent interactions. [34] A quintessential example is found in the work detailing the synthesis of a chain-like organic cation L·Cl2 from triethylenediamine and 1,2-bis(2-chloroethoxy)ethane, which then templates the formation of various metal-hybrid supramolecules. [34] The resulting structures can be zero-dimensional (mononuclear), one-dimensional chains, or higher-order networks, where the organic component often serves to balance charge and modify the structural properties of the inorganic framework. [34] [88]
For this comparison, "traditional catalysts" are categorized as follows:
The catalytic efficacy of ICS-derived nanocatalysts is evaluated against traditional systems across several key reactions. The data below summarize quantitative findings.
Table 1: Performance Comparison in Degradation and Synthesis Reactions
| Model Reaction | Catalyst Type | Specific Catalyst | Key Performance Metrics | Reference |
|---|---|---|---|---|
| Photocatalytic Degradation of Tetracycline | ICS-derived Nanocatalyst | {[(L)(Cu2I3)]·[CuI2]CH3CN}n (Compound 3) |
92.22% Degradation Efficiency (10 mg catalyst, pH=7); >86% efficiency after 4 cycles. [34] | |
| Traditional Heterogeneous Catalyst | Conventional TiO2-based photocatalysts | Typically lower degradation efficiency and significant activity loss over multiple cycles. [34] | ||
| Synthesis of N-Heterocycles (e.g., Pyrroles) | Magnetic ICS-inspired Nanocatalyst | Cu@imine/Fe3O4 Magnetic Nanoparticles (MNPs) | Excellent yields; Reusable 5-6 cycles; Solvent-free conditions. [91] | |
| Traditional Homogeneous Catalyst | Molecular Cu salts or acids | Difficult separation; Not easily recyclable; Often requires solvents. [91] | ||
| Carbonylation Reactions | Magnetic Nanoparticle-supported Catalyst | Transition Metals on MNPs (e.g., Pd, Ru) | High reaction rates & selectivity; Simple magnetic recovery; Mitigates waste. [90] | |
| Traditional Heterogeneous Catalyst | Supported transition metals (e.g., on Alumina) | Lower selectivity; Active site leaching; Complex filtration recovery. [90] |
Table 2: Inherent Characteristics and Economic & Environmental Impact
| Characteristic | ICS-Derived Nanocatalysts | Traditional Homogeneous Catalysts | Traditional Heterogeneous Catalysts |
|---|---|---|---|
| Structural Order & Tunability | High (Precise, programmable assembly) [34] | Moderate (Limited to ligand design) | Low (Random active site distribution) |
| Active Site Accessibility | High (Well-defined, exposed sites) [34] | High (All atoms accessible) | Variable (Diffusion limitations) |
| Recovery & Reusability | High (Especially magnetic ICS-hybrids) [90] [91] | Very Low (Difficult separation) | High (Simple filtration) |
| Stability & Leaching Resistance | High (Stable coordination environment) [34] | Moderate ( ligand decomposition) | Moderate (Sintering, poisoning) |
| Catalyst Cost | Moderate to High (Synthesis complexity) | Low to Moderate | Low to High (e.g., Pt) |
| Environmental Impact | Lower (Reusable, less waste) [90] | Higher (Solvent use, metal waste) | Moderate (Energy-intensive preparation) |
| Scalability Potential | Developing (Room-temperature synthesis promising) [34] | High (Established processes) | High (Established processes) |
This protocol details the synthesis of {[(L)(Cu2I3)]·[CuI2]CH3CN}n (Compound 3) and its use in tetracycline degradation. [34]
Synthesis of the Organic Cation Template (L·Cl2):
L·Cl2 ligand as a solid. Purify further via recrystallization.Synthesis of ICS Compound 3:
L·Cl2 and Copper(I) Iodide (CuI) in a 1:2 molar ratio in acetonitrile.Photocatalytic Degradation of Tetracycline:
This protocol outlines the use of a magnetic nanocatalyst, Cu@imine/Fe3O4 MNPs, for the solvent-free synthesis of polysubstituted pyrroles. [91]
Catalyst Preparation (Cu@imine/Fe3O4 MNPs):
Multicomponent Synthesis of Pyrroles:
The following diagram illustrates the multi-level architecture of a typical ICS-derived nanocatalyst, from molecular precursors to the final functional material.
This workflow outlines the key steps for evaluating the catalytic performance and reusability of an ICS-derived nanocatalyst, particularly a magnetic one.
Table 3: Key Reagents and Materials for ICS Nanocatalyst Research
| Reagent/Material | Function/Application | Technical Notes |
|---|---|---|
| 1,4-Diazabicyclo[2.2.2]octane (DABCO) | A common bicyclic amine used to construct organic cation templates for ICS structures. [34] | Its rigid structure and exposed nitrogen atoms favor stable complex formation. |
| Metal Salts (e.g., CuI, HgIâ, CdIâ, CoClâ) | Source of inorganic metal clusters (e.g., [CuâIâ]â», [HgIâ]²â»). The metal and anion choice dictates cluster geometry. [34] |
Use high-purity (>99%) salts. Anion selection influences solubility and final structure. |
| Functionalized Magnetic Nanoparticles (e.g., FeâOâ@SiOâ) | Core-shell support for creating magnetically recoverable ICS-hybrid catalysts. [90] [91] | The silica shell prevents core oxidation and provides a surface for functional group grafting (e.g., -SOâH, imine). |
| Linker Molecules (e.g., 1,2-bis(2-chloroethoxy)ethane) | Used to synthesize complex organic cations that template the assembly of the inorganic cluster. [34] | Chain length and functional groups (e.g., -Cl) determine the geometry and charge of the final cation. |
| Solvents for Crystallization (e.g., Acetonitrile, DMF) | Medium for the slow volatilization synthesis of single crystals suitable for X-ray diffraction. [34] | Must be high-purity and anhydrous. Solvent choice can direct the crystallization of different polymorphs. |
The development of supramolecular inorganic compounds represents a frontier in biomedical science, offering novel mechanisms for therapeutic intervention. This whitepaper provides a comprehensive technical evaluation of inhaled corticosteroids (ICSs) as a benchmark for localised therapy, contrasting their efficacy and mechanisms with emerging supramolecular metal-based structures. Within the context of inorganic cluster supramolecular compound synthesis research, we analyse pharmacological profiles, experimental methodologies, and clinical translation pathways to establish rigorous evaluation frameworks for novel therapeutic agents. Our integrated analysis reveals that optimized pharmacokinetic properties and targeted delivery systems are critical determinants of therapeutic efficacy across diverse compound classes.
Supramolecular chemistry has enabled the rational design of complex molecular architectures with precise biological functionalities. While conventional chemotherapeutics have dominated oncology treatment for decades, their systemic toxicity profiles necessitate the development of targeted alternatives. Inhaled corticosteroids (ICSs) represent a successful paradigm of localised therapy with minimized systemic exposure, achieving pulmonary selectivity through optimized pharmacokinetic and pharmacodynamic properties [92]. Meanwhile, supramolecular coordination complexes (SCCs) and supramolecular organometallic complexes (SOCs) represent emerging classes of metal-based structures with promising biomedical applications [93].
The evaluation of these diverse therapeutic classes requires sophisticated in vitro and in vivo models that can elucidate their mechanisms of action, efficacy, and safety profiles. This technical guide establishes a standardized framework for comparative analysis, emphasizing the critical pharmacological principles that govern therapeutic efficacy across compound classes. By examining established ICS therapeutics alongside emerging supramolecular structures, we aim to provide researchers with robust methodologies for advancing novel inorganic clusters through the drug development pipeline.
ICSs exhibit their therapeutic effects through complex molecular interactions with glucocorticoid receptors (GRs). The genomic effects involve diffusion across cell membranes, binding to cytoplasmic GRs, translocation to the nucleus, and modulation of gene transcription through transactivation and transrepression mechanisms [94]. This process regulates the expression of anti-inflammatory proteins and inhibits inflammatory mediators. Additionally, ICSs demonstrate non-genomic effects through specific interactions with membrane-bound receptors, providing rapid therapeutic effects within minutes of administration [94].
The pulmonary targeting of ICSs is governed by critical pharmacokinetic properties including high receptor binding affinity, lipophilicity that enhances tissue retention, and rapid systemic clearance that minimizes off-target effects [92]. The structural optimization of ICS molecules has progressively enhanced their therapeutic indices through prolonged pulmonary residence times and reduced oropharyngeal deposition.
Table 1: Comparative Pharmacological Properties of Selected Inhaled Corticosteroids
| Corticosteroid | Receptor Binding Affinity | Oral Bioavailability (%) | Systemic Clearance | Volume of Distribution | Plasma Half-life |
|---|---|---|---|---|---|
| Beclomethasone dipropionate | 0.4 (relative) | 20-40 | High | High | Short |
| Budesonide | 1.0 (reference) | 11 | High | Moderate | Intermediate |
| Fluticasone propionate | 1.8-3.0 | <1-2 | High | Extensive | Prolonged |
| Ciclesonide (active metabolite) | 1.0-1.2 | <1 | High | Extensive | Prolonged |
| Mometasone furoate | 1.5-2.0 | <1 | High | Extensive | Prolonged |
Supramolecular metal-based structures represent a paradigm shift in therapeutic development, leveraging coordination-driven self-assembly to create discrete architectures with defined biological activities. Supramolecular coordination complexes (SCCs) are typically formed through the combination of metal ions (nodes) with multidentate organic ligands (linkers) [93]. These structures exhibit unique host-guest capabilities that enable molecular encapsulation and targeted delivery. The edge-directed and face-directed assembly approaches allow precise control over size, shape, and functionality [93].
Emerging research demonstrates that supramolecular organometallic complexes (SOCs) incorporate metal-carbon bonds that enhance stability and create static structures with prolonged biological residence times [93]. These compounds, particularly those featuring N-heterocyclic carbene (NHC) ligands, exhibit remarkable stability under physiological conditions, making them promising candidates for therapeutic applications. The modular nature of these assemblies enables incorporation of imaging moieties, targeting ligands, and stimulus-responsive elements for theranostic applications.
The therapeutic efficacy of ICSs has been extensively quantified through randomized controlled trials and real-world studies across diverse patient populations. In pediatric asthma management, direct comparison of different ICSs revealed significant differences in onset kinetics and symptom control. Fluticasone propionate (FP) demonstrated superior efficacy in forced expiratory volume (FEV1) improvement compared to ciclesonide (CIC) and budesonide (BUD), with CIC exhibiting faster onset (ET~50~ 1.23 weeks) compared to BUD (ET~50~ 2.97 weeks) [95].
In COPD management, combination therapy with long-acting β-agonists (LABAs) and ICSs demonstrated a significant reduction in moderate exacerbations (RR 0.84; 95% CI 0.74-0.96) without affecting severe exacerbations or all-cause mortality [96]. Importantly, long-term ICS administration moderates lung function decline in asthma patients, reducing annual FEV1 decline by 23 mL/year compared to non-users [97].
Table 2: Comparative Efficacy of ICSs in Respiratory Diseases
| Therapeutic Application | Efficacy Endpoint | Budesonide | Fluticasone | Ciclesonide | Beclomethasone |
|---|---|---|---|---|---|
| Pediatric Asthma (FEV1 improvement) | Maximum efficacy (% change) | Moderate | High | Moderate | Lower |
| Pediatric Asthma | Onset kinetics (ET~50~, weeks) | 2.97 | - | 1.23 | - |
| Pediatric Asthma | Symptom-free days (%) | Moderate | High | High | Lower |
| Asthma (Real-world study) | Exacerbation reduction | Comparable | Comparable | - | - |
| COPD with LABA | Moderate exacerbation reduction (RR) | 0.84 (class effect) | - | - | - |
| Long-term asthma | FEV1 decline reduction (mL/year) | 23 (class effect) | - | - | - |
While clinical data for supramolecular metal complexes remains emerging compared to established ICSs, preclinical studies demonstrate promising therapeutic potential. The unique molecular recognition properties of metallacages enable host-guest chemistry that can encapsulate conventional chemotherapeutics, enhancing their solubility and targeted delivery [93]. Specific SCCs exhibit inherent anticancer activity through novel mechanisms of action distinct from traditional chemotherapeutics, potentially overcoming resistance pathways.
The integration of therapeutic and diagnostic capabilities (theranostics) within single supramolecular structures represents a significant advancement beyond conventional ICS capabilities. These systems enable real-time monitoring of drug distribution and therapeutic response, facilitating personalized treatment approaches. The robust and modular composition of metal-based supramolecular structures allows for systematic optimization of efficacy parameters through controlled variation of metal nodes and organic linkers.
Primary Human Airway Epithelial Cell Culture Protocol:
Receptor Binding Affinity Assay:
Pulmonary Pharmacokinetic Study Design:
Pharmacodynamic Response Modeling:
Table 3: Essential Research Reagents for ICS and Supramolecular Compound Evaluation
| Reagent/Category | Specific Examples | Research Function | Technical Considerations |
|---|---|---|---|
| Cell Culture Models | Primary human airway epithelial cells, BEAS-2B, A549, THP-1 | In vitro efficacy screening | Air-liquid interface culture mimics in vivo conditions |
| Inflammatory Stimuli | TNF-α, IL-1β, LPS, ovalbumin | Induction of inflammatory pathways | Concentration and timing critical for reproducible responses |
| Molecular Biology Assays | ELISA kits, RT-qPCR reagents, Western blot systems | Quantification of inflammatory mediators | Validate antibody specificity for target analytes |
| Animal Models | Ovalbumin-sensitized mice, house dust mite models | In vivo efficacy and safety assessment | Strain-specific differences in immune responses |
| Analytical Standards | ³H- or ¹â´C-labeled compounds, certified reference standards | PK study quantification | Ensure radiochemical purity and stability |
| Drug Delivery Devices | Microsprayers, vibrating mesh nebulizers, DPIs, MDIs | Controlled compound administration | Device characteristics affect pulmonary deposition |
| Receptor Binding Assays | Recombinant GR, radiolabeled ligands | Mechanism of action studies | Control for non-specific binding in calculations |
The comparative evaluation of ICSs and supramolecular metal-based structures reveals several convergent principles for optimal therapeutic design. Both therapeutic classes benefit from enhanced local retention at the target site, whether pulmonary tissue for ICSs or tumor microenvironments for supramolecular complexes. The structural modularity of SCCs mirrors the progressive optimization of ICS molecules, enabling fine-tuning of therapeutic indices through systematic variation of component parts.
Future research directions should focus on advancing the theranostic capabilities of supramolecular structures, integrating real-time monitoring with therapeutic intervention. The development of stimulus-responsive systems that activate specifically in disease microenvironments represents a promising approach to enhance selectivity. Additionally, the application of quantitative systems pharmacology models to supramolecular complexes will facilitate translational prediction from in vitro data to clinical outcomes.
The integration of supramolecular chemistry principles with established pharmacological evaluation frameworks presents unprecedented opportunities for innovation in therapeutic development. By leveraging the quantitative insights from ICS research and the structural diversity of metal-based assemblies, researchers can accelerate the development of next-generation therapeutics with enhanced efficacy and safety profiles.
This technical evaluation establishes rigorous frameworks for comparative assessment of ICSs and emerging supramolecular therapeutics. The quantitative efficacy data, standardized experimental protocols, and essential research tools provide researchers with comprehensive methodologies for advancing novel compounds through the development pipeline. The integration of pharmacological principles with supramolecular design represents a transformative approach for creating targeted therapies with optimized therapeutic indices across diverse disease contexts.
Structure-Activity Relationship (SAR) analysis represents a fundamental methodology in molecular design that systematically investigates how chemical structure modifications influence biological activity or material properties. In the specific context of inorganic cluster supramolecular compounds, SAR studies enable researchers to decipher how alterations to molecular components affect the resulting non-covalent interactions, self-assembly behavior, and ultimately, the functional performance of these sophisticated systems. Supramolecular chemistry, defined as "chemistry beyond the molecule," focuses on chemical systems composed of discrete numbers of molecules organized through non-covalent interactions including hydrogen bonding, metal coordination, hydrophobic forces, van der Waals forces, ÏâÏ interactions, and electrostatic effects [10]. The primary goal of SAR analysis in this field is to establish predictive relationships that can guide the rational design of supramolecular materials with tailored properties for applications ranging from drug delivery to nonlinear optics and molecular sensing.
The philosophical roots of supramolecular chemistry trace back to 1894 with Emil Fischer's "lock and key" hypothesis of enzyme-substrate interactions, establishing the fundamental principles of molecular recognition [10]. The field gained substantial momentum with the discovery of crown ethers by Charles J. Pedersen in 1967, followed by contributions from Jean-Marie Lehn and Donald J. Cram, who collectively earned the 1987 Nobel Prize in Chemistry for "development and use of molecules with structure-specific interactions of high selectivity" [10]. Today, supramolecular chemistry is experiencing a translational shift toward applying fundamental understanding to commercially viable products, moving innovation from laboratories into the commercial marketplace [47]. Within this framework, SAR analysis serves as the critical bridge connecting molecular-level structural insights with macroscopic functional performance.
Supramolecular systems derive their structure and function from directional non-covalent interactions that operate in concert to stabilize complex architectures. These interactions exhibit varying strength ranges and geometric preferences that profoundly influence supramolecular organization:
The design of inorganic cluster supramolecular compounds relies on strategic selection of structural units with defined geometries and interaction potentials:
The interplay between these building blocks and their interaction profiles forms the structural basis for SAR investigations in supramolecular systems.
Computational methods provide powerful tools for predicting structure-property relationships before embarking on complex synthetic pathways. Several specialized approaches have been developed specifically for supramolecular systems:
Molecular Modeling and Density Functional Theory (DFT) Calculations DFT has emerged as an indispensable computational method for understanding the relationship between geometric and electronic structures in supramolecular systems [98]. For chalcone-based systems with donor-Ï-acceptor (D-Ï-A) configurations, DFT calculations successfully predicted enhanced first (β) and second (γ) hyperpolarizabilities, guiding the synthesis of compounds with improved nonlinear optical properties [98]. These calculations enable researchers to model intramolecular charge transfer (ICT) characteristics and predict how structural modifications will influence electronic properties and supramolecular organization.
R-group Analysis and Matched Pair Analysis R-group analysis systematically explores the effects of substituent variations on supramolecular properties, enabling researchers to identify optimal functional groups for specific applications [99]. This approach is particularly valuable when working with core scaffold structures such as porphyrins, phthalocyanines, or inorganic clusters where peripheral substitutions dramatically influence assembly behavior. Matched pair analysis extends this methodology by identifying molecular pairs that differ by a single structural transformation, allowing researchers to quantify the specific contribution of individual modifications to supramolecular properties [99].
Activity Landscape Analysis Activity landscape analysis creates networks of similar compounds to identify "activity cliffs"âregions where small structural changes produce dramatic property variations [99]. In supramolecular chemistry, this approach can reveal critical structural elements that control self-assembly pathways or molecular recognition events. Similarly, this method can identify regions of "flat SAR" where extensive modifications produce minimal property changes, highlighting robust structural motifs that maintain function across diverse chemical contexts [99].
Table 1: Computational SAR Methods for Supramolecular Systems
| Method | Key Function | Supramolecular Application | Tools/Software |
|---|---|---|---|
| DFT Calculations | Predict electronic properties and non-covalent interactions | Modeling charge transfer in D-Ï-A systems; host-guest binding energies | Gaussian, ORCA |
| R-group Analysis | Systematically vary substituents | Optimize peripheral groups on macrocycles or cluster ligands | DataWarrior [99], StarDrop [99] |
| Matched Pair Analysis | Isolate effects of single modifications | Quantify contribution of specific functional groups to binding | StarDrop [99], NextMove Software |
| Activity Landscape Analysis | Identify critical structural regions | Map self-assembly behavior across structural space | DataWarrior [100], StarDrop [99] |
| Molecular Dynamics | Simulate assembly processes | Model solvation effects and pathway complexity | GROMACS, AMBER |
Experimental validation represents a critical component of SAR studies, providing essential data to correlate with computational predictions. Several advanced characterization methods are particularly relevant to inorganic cluster supramolecular systems:
X-ray Crystallography Single-crystal X-ray diffraction provides atomic-level resolution of supramolecular structures, revealing precise geometric parameters and interaction motifs [98]. For example, in the study of methoxy-substituted thiophene chalcones, X-ray crystallography confirmed molecular planarity and revealed specific intermolecular contacts responsible for crystal packing [98]. This structural information directly correlates with observed nonlinear optical properties, enabling robust SAR conclusions.
Spectroscopic Analysis Multinuclear NMR spectroscopy (particularly (^1)H NMR titration experiments) quantifies host-guest binding constants and reveals interaction stoichiometries in solution. Fourier-transform infrared (FTIR) spectroscopy identifies characteristic vibrational modes that change upon complexation or assembly, providing evidence for specific interaction types [98]. Mass spectrometry, especially electrospray ionization (ESI-MS), confirms the formation of discrete supramolecular complexes and can detect species that are labile in solution.
Thermal and Morphological Characterization Thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) assess the thermal stability and phase behavior of supramolecular materials, properties that directly influence their practical applications. Scanning electron microscopy (SEM) and atomic force microscopy (AFM) visualize the nanoscale and microscale morphologies resulting from self-assembly processes, connecting molecular structure to macroscopic organization.
Table 2: Experimental Techniques for Supramolecular SAR Studies
| Technique | Structural Information | Functional Properties | SAR Relevance |
|---|---|---|---|
| X-ray Crystallography | Atomic coordinates, bond lengths/angles, packing motifs | Solid-state organization | Direct structure-property correlations |
| NMR Spectroscopy | Solution-state structure, binding constants, dynamics | Host-guest interactions, molecular recognition | Quantify interaction strength |
| UV-Vis/NIR Spectroscopy | Electronic transitions, charge transfer bands | Optical properties, sensing capabilities | Relate structure to electronic properties |
| Fluorescence Spectroscopy | Excited state behavior, energy transfer | Photophysical properties, signaling | Design functional materials |
| Isothermal Titration Calorimetry | Binding thermodynamics (ÎH, ÎS) | Affinity, selectivity | Understand interaction forces |
| Atomic Force Microscopy | Nanoscale morphology, mechanical properties | Surface organization, material behavior | Connect molecular to macroscopic |
A recent investigation of (E)-1-(5-bromothiophen-2-yl)-3-(3,4-dimethoxy-phenyl)-prop-2-en-1-one (DBR-19) provides an exemplary case study of SAR principles applied to supramolecular material design [98]. This research demonstrates a comprehensive approach to understanding how specific structural elements influence macroscopic properties in a chalcone-based system with potential nonlinear optical (NLO) applications.
The target compound DBR-19 was designed with a donor-Ï-donor (D-Ï-D) configuration featuring a bromine substituent as an electron-donating group on the thiophene ring and dimethoxy groups at the 3rd and 4th positions on the phenyl ring [98]. The synthesis employed a straightforward Claisen-Schmidt condensation reaction between 2-acetyl-5-bromo-thiophene and 3,4-dimethoxy-benzaldehyde (veratraldehyde) using lithium hydroxide monohydrate as a base catalyst in ethanol [98]. This synthetic approach highlights the importance of selecting appropriate reaction conditions to obtain the desired supramolecular building block with high purity and yield.
X-ray crystallographic analysis revealed that DBR-19 crystallizes in the monoclinic system with space group P2(_1)/c, displaying significant molecular planarity across the entire conjugated system [98]. This structural feature facilitates effective intramolecular charge transfer (ICT), a crucial requirement for enhanced NLO properties. The study identified specific intermolecular interactions, including C-H···O hydrogen bonds and Ï-Ï stacking, that stabilize the crystal packing [98]. Hirshfeld surface analysis quantified the contribution of different intermolecular contacts, with H···H (36.2%), O···H/H···O (19.7%), and S···H/H···S (11.1%) interactions dominating the crystal environment [98].
DFT calculations performed at the B3LYP/6-311++G(d,p) level provided complementary electronic structure information, showing that the bromine and methoxy substituents synergistically enhance intramolecular charge transfer characteristics [98]. The theoretical hyperpolarizability (β) value of 12.943 à 10(^{-30}) esu confirmed significant NLO potential, which was experimentally verified through second-harmonic generation (SHG) measurements that revealed an efficiency approximately 1.8 times that of standard potassium dihydrogen phosphate (KDP) [98]. This combined experimental and theoretical approach established clear relationships between specific structural features (planarity, donor groups, conjugation length) and the resulting NLO performance.
SAR Relationships in NLO Materials
Protocol: Claisen-Schmidt Condensation for Chalcone Synthesis [98]
Objective: Prepare conjugated organic building blocks with D-Ï-A or D-Ï-D configurations for supramolecular assembly and NLO applications.
Materials:
Procedure:
Protocol: Single Crystal Growth and X-ray Diffraction Analysis [98]
Objective: Obtain high-quality single crystals for precise structural determination of supramolecular organization.
Materials:
Procedure:
Protocol: UV-Vis and Fluorescence Spectroscopy for Electronic Properties
Objective: Characterize electronic transitions and intramolecular charge transfer behavior in supramolecular systems.
Materials:
Procedure:
Table 3: Essential Research Reagents for Supramolecular SAR Studies
| Reagent/Material | Function | Application Example | Supplier Examples |
|---|---|---|---|
| Functionalized Building Blocks | Core structural elements for assembly | Thiophene chalcones, porphyrins, crown ethers | Sigma-Aldrich, TCI Chemicals |
| Metal Salts | Coordination centers for metallo-supramolecular systems | Zn(II), Cu(II), Fe(II/III), Ln(III) salts | Strem Chemicals, Alfa Aesar |
| Spectroscopic Solvents | Medium for solution-state studies | Deuterated solvents for NMR, anhydrous solvents | Cambridge Isotopes, Sigma-Aldrich |
| Crystallization Solvents | Growth of single crystals for XRD | HPLC grade MeOH, EtOH, CHâCN, mixed solvent systems | Fisher Scientific, Merck |
| Computational Software | Molecular modeling and property prediction | Gaussian, ORCA, DataWarrior [100] | Academic/commercial licenses |
| Chromatography Materials | Purification of synthetic compounds | Silica gel, TLC plates, HPLC columns | Silicycle, Waters, Agilent |
| Spectroscopic Standards | Calibration and quantification | NMR reference compounds, quantum yield standards | Sigma-Aldrich, commercial sources |
| Specialized Glassware | Controlled reactions and crystallization | Schlenk ware, crystallization tubes, NMR tubes | Chemglass, Wilmad-LabGlass |
The field of SAR analysis for supramolecular material design is rapidly evolving, with several emerging trends poised to significantly advance research capabilities. The integration of automated synthesis and high-throughput screening platforms represents a particularly promising direction, enabling researchers to efficiently explore vast chemical spaces and identify optimal structural combinations [47] [78]. These automated approaches are especially valuable for investigating complex, multi-component supramolecular systems where traditional one-at-a-time experimentation would be prohibitively time-consuming and resource-intensive.
Another significant development involves the creation of large-scale datasets and benchmark collections specifically tailored for supramolecular systems. Following the example established by initiatives like SARDet-100K in synthetic aperture radar [101], the supramolecular community would benefit greatly from standardized datasets that enable direct comparison of different structural families and assembly motifs. Such resources would facilitate the application of machine learning algorithms to predict supramolecular behavior and guide material design.
The ongoing translation of supramolecular chemistry from fundamental research to commercial applications further highlights the practical importance of robust SAR methodologies [47]. As researchers increasingly focus on developing "real-world" applications for supramolecular systems, the ability to precisely correlate structural features with functional performance becomes increasingly critical. This translational emphasis is evident in diverse applications ranging from pharmaceutical formulations utilizing cyclodextrin complexes [47] to advanced materials with tailored electronic, optical, or mechanical properties [98].
In conclusion, SAR analysis provides an essential framework for advancing the design of inorganic cluster supramolecular compounds. By systematically correlating structural modifications with changes in supramolecular behavior and functional properties, researchers can move beyond serendipitous discovery toward rational material design. The continued refinement of computational and experimental SAR methodologies, coupled with emerging technologies in automation and data science, promises to significantly accelerate the development of next-generation supramolecular materials with precisely tailored properties for advanced technological applications.
The development of new chemical entities, particularly within innovative fields like inorganic cluster supramolecular compound synthesis, necessitates rigorous safety evaluation to determine their potential for biomedical application. Biocompatibility and toxicity profiling form the cornerstone of this assessment, providing critical data on how these materials interact with biological systems. Comprehensive toxicity screening is essential for the drug development process, helping researchers characterize potential adverse effects and establish preliminary safety margins to determine if a compound is safe for human use [102] [103]. For inorganic supramolecular compounds, which often possess unique structural and chemical properties, this evaluation is particularly crucial as their complex architecture may introduce novel interaction mechanisms with biological systems.
The historical foundation of toxicology, famously summarized by Paracelsus (1493-1541) who stated "All substances are poisons; there is none which is not a poison. The right dose differentiates a poison and a remedy," remains fundamentally relevant to modern compound safety assessment [103]. This principle is especially pertinent when evaluating synthetic inorganic compounds where therapeutic potential and toxicity must be carefully balanced. The preclinical toxicity testing on various biological systems reveals the species-, organ- and dose-specific toxic effects of an investigational product, providing essential data for calculating the "No Observed Adverse Effect Level" (NOAEL) necessary to initiate clinical evaluation [103]. Within the specific context of inorganic supramolecular chemistry research, where compounds may be designed for applications ranging from catalytic functions to biomaterial integration, understanding these toxicological parameters is indispensable for guiding synthetic strategies toward biologically compatible architectures.
In vitro cytotoxicity evaluation serves as the initial screening platform in biocompatibility assessment, providing rapid, cost-effective data on the intrinsic toxicity of materials before progressing to complex in vivo studies. These assays utilize isolated cells to evaluate the toxicity or irritancy potential of materials and chemicals, offering an excellent way to screen materials prior to in vivo tests [104]. For inorganic supramolecular compounds, which may contain metal centers and ligand arrangements with unknown cellular interactions, these preliminary screens are invaluable for identifying overt cytotoxicity and guiding structural refinement.
The three qualitative cytotoxicity tests commonly used for material evaluation include the Direct Contact, Agar Diffusion, and MEM Elution assays, each with specific applications depending on material properties [104]. The Direct Contact procedure is recommended for low-density materials, where a piece of test material is placed directly onto cells growing on culture medium, allowing leachable chemicals to diffuse into the culture medium and contact the cell layer. Reactivity is indicated by malformation, degeneration and lysis of cells around the test material. The Agar Diffusion assay is appropriate for high-density materials, utilizing a thin layer of nutrient-supplemented agar placed over cultured cells, with the test material positioned on top. The MEM Elution assay uses different extracting media and extraction conditions to test devices according to actual use conditions or to exaggerate those conditions, with extracts transferred onto cell layers and incubated before microscopic examination for malformation, degeneration and lysis [104].
For inorganic supramolecular compounds, where solubility and ion release kinetics may significantly influence biological interactions, the MEM Elution method is particularly valuable as it allows evaluation of extractable components under varied physiological conditions.
Recent regulatory guidelines (ANSI/AAMI/ISO 10993-5:2009) on biocompatibility for devices state that while qualitative cytotoxicity tests are appropriate for screening purposes, quantitative evaluation is preferable [104]. The MTT cytotoxicity assay has emerged as a fundamental quantitative method, accurately quantifying as few as 950 cells through a colorimetric approach that measures the reduction of yellow 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide by mitochondrial succinate dehydrogenase [104]. This assay provides significant advantages for evaluating inorganic compounds, including its quantitative nature, applicability to either extracts or direct contact scenarios, and reduced analyst interpretation variability. Additionally, the method facilitates high-throughput screening when performed on 96-well microplates, enabling efficient assessment of multiple compound variants or concentrations.
Table 1: Standardized In Vitro Cytotoxicity Assays for Biocompatibility Screening
| Assay Method | Principle of Detection | Application Context | Key Advantages | Research Considerations |
|---|---|---|---|---|
| Direct Contact | Cellular morphology changes near test material | Low-density materials and direct interfacial interactions | Direct simulation of device-tissue interface; no extraction variability | Limited for soluble compounds; requires physical material sample |
| Agar Diffusion | Diffusion through agar barrier to cell layer | High-density materials and elastomeric compounds | Barrier protects cells from physical damage; detects diffusible toxins | Semi-quantitative; diffusion kinetics influenced by material properties |
| MEM Elution | Extraction in physiological media followed by cell exposure | Evaluation of leachables and extractables | Adaptable extraction conditions; mimics physiological exposure | Extraction parameters critical; may not reflect particle effects |
| MTT Assay | Mitochondrial dehydrogenase activity in living cells | Quantitative cytotoxicity screening | High sensitivity; objective quantification; adaptable to high-throughput | Endpoint measurement only; may underestimate early cellular damage |
The application of these in vitro methods to inorganic supramolecular compounds is exemplified in research on ochratoxin α amide, a degradation product of ochratoxin A, where cytotoxicity screening using immortalized human kidney epithelial (IHKE) cells demonstrated no cytotoxicity up to concentrations of 50 μM, suggesting a detoxification process had occurred through structural modification [105]. This approach highlights how systematic in vitro evaluation can guide understanding of structure-activity relationships in complex molecular systems.
Materials Required:
Procedure:
For inorganic supramolecular compounds, particular attention should be paid to appropriate solvent systems that maintain compound stability, potential interference with assay detection systems, and relevant exposure durations that reflect anticipated application contexts.
Figure 1: MTT Cytotoxicity Assay Workflow. This diagram outlines the key procedural steps for conducting quantitative cytotoxicity assessment using the MTT assay, a fundamental method for evaluating compound toxicity in vitro.
In vivo toxicological evaluation represents a critical translational step in biocompatibility assessment, providing comprehensive data on systemic effects that cannot be fully replicated in simplified in vitro systems. These studies are designed to assess the onset, severity, duration, dose dependency and irreversibility of toxic effects in whole organisms, capturing complex physiological interactions and metabolic processes [102]. For inorganic supramolecular compounds, whose behavior in biological environments may involve dissociation, metal ion release, or novel distribution patterns, in vivo assessment is particularly vital for establishing realistic safety profiles.
Acute toxicity testing is carried out to determine the effects of a single dose administration on a particular animal species, typically using two different animal species (one rodent and one non-rodent) [103]. During acute toxicological evaluation, the investigational product is administered at different dose levels, and effects are observed for 14 days, with all mortalities and morphological, biochemical, pathological, and histological changes recorded and investigated. Historically, the 50% lethal dose (LDâ â) test was used as a standard indicator of acute toxicity, but due to animal welfare concerns and methodological limitations, modified approaches including the Fixed Dose Procedure (FDP), Acute Toxic Category (ATC) method, and Up-and-Down (UDP) method have been developed [103].
The Up-and-Down testing approach is particularly recommended by various regulatory agencies as it significantly reduces the number of vertebrate animals required for testing. This sequential method involves dosing single animals at 48-hour intervals, beginning with a dose less than the best-estimate LDâ â [103]. If the animal survives, testing continues with a higher dose; if the animal dies, testing proceeds with a lower dose using another animal of the same sex. This method is particularly efficient for establishing toxicity thresholds for novel inorganic compounds where preliminary toxicity data may be limited.
Repeated dose toxicity testing extends exposure duration to identify potential adverse effects from longer-term or multiple exposures, typically conducted over periods up to 28 days or extended to 90 days for subchronic evaluation [103]. These studies are required for all permanent devices and should be considered for materials with prolonged contact with internal tissues [104]. The study design typically involves daily administration of the test substance through the intended clinical route, with comprehensive monitoring of clinical signs, body weight, food and water consumption, hematological and biochemical parameters, and detailed histopathological examination of tissues at study termination.
For inorganic supramolecular compounds, which may demonstrate accumulation in specific tissues or progressive structural modifications in biological environments, these longer-term studies are essential for identifying potential chronic toxicity concerns. An example of this approach is demonstrated in a study of Syzygium aqueum leaf extract, where subchronic toxicity was evaluated through oral administration at doses of 50, 100, and 200 mg/kg per day for 28 days to male Sprague-Dawley rats, with comprehensive assessment of food and water intake, body weight, organ weights, biochemical parameters and histopathological observations [106]. The results demonstrated no significant differences between control and treated rats across all parameters, indicating an absence of subchronic toxicity at the tested doses.
Materials Required:
Procedure:
Table 2: Standard In Vivo Toxicity Study Designs and Parameters
| Study Type | Standard Duration | Typical Species | Key Assessment Parameters | Regulatory Guidelines |
|---|---|---|---|---|
| Acute Toxicity | Single dose + 14-day observation | Rat, mouse (one rodent and one non-rodent) | Mortality, clinical signs, gross pathology | OECD 420, 423, 425 |
| Repeated Dose 28-Day | 28 days daily dosing | Rat, mouse | Clinical observations, body weight, food/water consumption, hematology, clinical chemistry, gross and histopathology | OECD 407 |
| Subchronic Toxicity | 90 days daily dosing | Rat, dog, non-human primate | Comprehensive clinical, hematological, biochemical, and histopathological evaluation; often includes toxicokinetics | OECD 413 |
| Skin Sensitization | Induction and challenge phases over 21-28 days | Guinea pig, mouse | Erythema, edema, lymphocyte proliferation in Local Lymph Node Assay (LLNA) | OECD 406, 429 |
Figure 2: In Vivo Toxicity Study Workflow. This diagram illustrates the key stages in conducting in vivo toxicity assessments, from study design through to data interpretation and reporting.
Emerging biosensor technologies offer transformative potential for biocompatibility assessment, enabling real-time, sensitive detection of biological responses to material exposure. Biosensors are analytical devices that utilize biological recognition elements such as enzymes, nucleic acids, antibodies, proteins, and peptides to detect target analytes, converting the result of molecular recognition into measurable electrical signals [108] [109]. For inorganic supramolecular compounds, whose biological effects may involve specific molecular interactions or subtle physiological changes, biosensors provide sophisticated tools for mechanistic toxicity evaluation beyond traditional endpoint analyses.
The fundamental architecture of biosensors comprises two main components: a bio-recognition element that identifies the target analyte, and a transducer that converts the molecular recognition event into an electrical signal [108]. Recent advances have integrated nanotechnology with biosensing platforms, creating devices with enhanced performance characteristics including improved sensitivity, specificity, and miniaturization potential [110]. Electrochemical biosensors represent a particularly promising category for toxicity assessment, employing electrodes that convert chemical signals into electrical outputs, enabling detection of various biomolecules in biological systems including glucose, cholesterol, uric acid, lactate, DNA, hemoglobin, and blood ketones [108].
The application of biosensor-integrated drug delivery systems represents a particularly advanced approach to toxicity management, creating closed-loop systems that can both monitor physiological changes and initiate therapeutic responses [108] [109]. These systems typically consist of a monitoring component that senses surrounding conditions and an actuator component with capability to trigger drug release, allowing intervention when biomarker levels exceed threshold values while maintaining normal function during homeostatic conditions [109]. While originally developed for chronic disease management, this conceptual framework has significant implications for managing potential adverse effects of bioactive materials, including inorganic supramolecular compounds with therapeutic applications.
Successful biocompatibility and toxicity profiling requires access to specialized reagents, materials, and analytical systems. The following table summarizes key resources essential for conducting comprehensive safety assessment of inorganic supramolecular compounds.
Table 3: Essential Research Reagents and Materials for Biocompatibility Assessment
| Category/Item | Specification and Selection Criteria | Primary Function in Assessment | Application Notes for Inorganic Compounds |
|---|---|---|---|
| Cell Culture Systems | Immortalized human kidney epithelial (IHKE) cells, 3T3-L1 fibroblasts, primary cell isolates | In vitro cytotoxicity screening; mechanistic toxicity studies | Select cell lines relevant to exposure route; consider specialized cells for metal metabolism (hepatocytes) |
| MTT Reagent | 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide, â¥97.5% purity | Quantitative assessment of cell viability through mitochondrial function | Validate absence of interference with test compounds; confirm solubility characteristics |
| Animal Models | Sprague-Dawley rats, CD-1 mice, species appropriate for toxicokinetics | In vivo safety and toxicity profiling | Consider specialized models for organ-specific accumulation assessment |
| Hematology Analyzer | Automated systems for complete blood count with differential | Evaluation of hematological toxicity and immune response | Establish baseline ranges for selected species; monitor for metal-induced hematological changes |
| Clinical Chemistry Analyzers | Platforms for measuring albumin, ALT, ALP, amylase, AST, BUN, calcium, cholesterol, creatinine, glucose, phosphate, total protein | Assessment of organ function and damage | Include metal-specific biomarkers when appropriate (e.g., metallothioneins) |
| Histopathology Equipment | Tissue processor, microtome, staining systems, light microscopy | Morphological evaluation of tissue damage at cellular level | Develop specialized staining protocols for inorganic compound detection |
| Biosensor Platforms | Electrochemical, optical, or thermal transducers with appropriate biorecognition elements | Real-time monitoring of biomarker changes | Customize recognition elements for specific inorganic compound interactions |
Biocompatibility and toxicity profiling represents an essential component in the development pathway of inorganic supramolecular compounds, providing critical safety data that guides compound selection, optimization, and eventual application. The multidisciplinary approach encompassing in vitro cytotoxicity screening, comprehensive in vivo safety assessment, and emerging biosensing technologies creates a robust framework for evaluating biological interactions of novel materials. For researchers in inorganic supramolecular chemistry, understanding and implementing these assessment strategies is crucial for translating synthetic achievements into biologically compatible applications with therapeutic potential. The standardized protocols and methodological frameworks presented in this guide provide a foundation for rigorous safety evaluation, enabling the development of innovative inorganic compounds with optimized biocompatibility profiles.
Inorganic Cluster Supramolecular compounds (ICSs) represent a advanced class of materials where molecular-level control converges with macroscopic functionality. These systems are characterized by the deliberate assembly of inorganic nanobuilding blocks (NBBs) and organic components through specific, reversible non-covalent interactions [111]. The foundational principle of supramolecular chemistryâgoverning the construction of complex structures from molecular components via interactions such as hydrogen bonding, metal coordination, and ÏâÏ stackingâprovides the toolbox for ICSs development [10]. This approach enables researchers to transcend the limitations of traditional covalent synthesis, offering a pathway to create materials with precisely engineered properties. The intrinsic reversibility of supramolecular bonds introduces dynamic, adaptive characteristics not found in conventional polymers or ceramics, paving the way for smart, responsive material systems [111].
The significance of ICSs emerges most clearly when contrasted with conventional materials. Traditional material design often involves trade-offs: enhancing mechanical strength may compromise processability, or incorporating functional properties can lead to structural instability. ICSs overcome these limitations through their inherent hierarchical organization. By employing well-defined inorganic clusters as primary building unitsâsuch as titanium oxo-clusters or molybdenum halide clustersâand directing their assembly via supramolecular directives, researchers can create organic-inorganic hybrids where both components contribute synergistically to the material's final properties [111] [112]. This paradigm shift from "making molecules" to "directing molecular assemblies" opens unprecedented opportunities for creating multifunctional systems with applications ranging from drug delivery to adaptive coatings and energy conversion.
Supramolecular chemistry provides the theoretical foundation for ICSs design, emphasizing molecular recognition and self-assembly as core principles [10]. Unlike conventional covalent materials, supramolecular systems maintain their integrity through precisely orchestrated non-covalent interactions including hydrogen bonding, metal coordination, hydrophobic forces, van der Waals forces, ÏâÏ interactions, and electrostatic effects [10] [113]. The strength of these individual interactions is relatively weak compared to covalent bonds (typically 2-300 kJ molâ»Â¹ versus 150-450 kJ molâ»Â¹), but when employed cooperatively and in sufficient numbers, they produce stable, well-defined architectures [113].
Several key concepts distinguish the supramolecular approach to material design:
Molecular Self-Assembly: The autonomous organization of components into structured aggregates without external guidance, driven by the complementarity of molecular recognition sites [10]. This process can be further classified into intermolecular self-assembly (forming supramolecular aggregates) and intramolecular self-assembly (folding), both crucial for ICSs construction.
Host-Guest Chemistry: The specific binding of complementary molecular components through non-covalent interactions, enabling the creation of complex architectures with defined cavities and functions [10]. This principle is exemplified in crown ether-metal ion complexes and cucurbituril-based systems.
Dynamic Covalent Chemistry: While supramolecular structures primarily rely on non-covalent bonds, the strategic implementation of reversible covalent bonds (such as disulfide formation) can introduce additional control over material assembly and responsiveness [113].
The historical development of supramolecular chemistry reveals its growing influence on material science. Beginning with Johannes Diderik van der Waals' postulation of intermolecular forces in 1873 and Hermann Emil Fischer's "lock and key" model of molecular recognition in 1894, the field matured through seminal discoveries including Charles J. Pedersen's crown ethers (1967), Jean-Marie Lehn's cryptands (1969), and Donald J. Cram's host-guest complexes [10]. The recognition of these pioneers with the 1987 Nobel Prize in Chemistry, followed by the 2016 Nobel Prize for molecular machines awarded to Sauvage, Stoddart, and Feringa, underscores the transformative potential of controlling matter at the supramolecular level [10].
Modularity represents a foundational advantage of ICSs, enabling researchers to systematically construct complex materials from discrete, well-defined building blocks. This approach employs a "molecular toolkit" where inorganic clusters and organic linkers with specific geometries and interaction profiles can be combined in predictable ways to generate targeted architectures [111] [6]. The modular design paradigm fundamentally distinguishes ICSs from conventional materials, where properties typically emerge from homogeneous molecular compositions rather than precisely orchestrated heterocomponent assemblies.
The inorganic component of ICSs provides structural definition and often introduces electronic, optical, or catalytic functionality. These clusters serve as molecular platforms that can be synthetically modified to present specific supramolecular interaction sites:
Titanium Oxo-Clusters: As demonstrated in pioneering work on hybrid dynamers, the TiââOââ(OEt)ââ cluster can be functionally tailored through post-synthetic modification, exchanging terminal ethoxy groups with hydrogen-bond accepting ligands like 1-(2-hydroxyethyl)-2-pyrrolidone [111]. This precise functionalization introduces molecular recognition capability while preserving the integrity of the inorganic core, as verified by ¹â·O NMR spectroscopy [111].
Molybdenum Halide Clusters: The photophysically active [{MoâIâ}(OOC-Câ HâN)â] cluster exemplifies how inorganic NBBs can be designed for supramolecular assembly, presenting six pyridine donor groups capable of coordinating to metalloporphyrin acceptors [112]. This modular design enables the construction of predictable cluster-porphyrin architectures (CPn, where n = 2, 6) with defined stoichiometries and geometries confirmed by single-crystal X-ray diffraction [112].
Zirconium-Based Clusters: Advanced Zrâ(μââO)â(μââOH)â nodes in MOF-type structures provide sterically open coordination sites that can support catalytic metal centers, demonstrating how cluster geometry dictates functional capability [114].
The organic partners in ICSs assembly introduce structural directionality and reversibility through programmed molecular recognition:
Telechelic Polymers: End-functionalized polymers like thiourea-terminated polydimethylsiloxane (PDMS) serve as flexible spacers that connect rigid inorganic clusters through specific hydrogen-bonding interactions [111]. The choice of terminal groups (e.g., thiourea as strong hydrogen bond donors) dictates the association strength and selectivity with complementary cluster surfaces.
Metalloporphyrins: As evidenced in molybdenum cluster-zinc porphyrin systems, metalloporphyrins function as multidentate organic acceptors that coordinate to pyridine-decorated clusters, forming discrete hybrid architectures with defined stoichiometries [112].
Supramolecular Synthons: The strategic pairing of complementary interaction groups directs ICSs assembly. The hydrogen-bonding duo of pyrrolidone acceptors (on clusters) and thiourea donors (on polymers) exemplifies a highly selective synthon that promotes hetero-association while minimizing competing homo-association [111].
Table 1: Modular Components in ICSs Design
| Component Type | Specific Examples | Key Features | Role in ICSs Assembly |
|---|---|---|---|
| Inorganic NBBs | TiââOââ(OEt)ââ with pyrrolidone ligands | ⢠Preserved oxo-core structure⢠Four hydrogen bond acceptors⢠Photochromic properties | Structural node with molecular recognition capability |
| [{MoâIâ}(OOC-Câ HâN)â] | ⢠Six pyridine coordination sites⢠Notable photophysical properties⢠Anionic character | Photophysically active module for coordination assemblies | |
| Organic Partners | Thiourea-terminated PDMS | ⢠Strong H-bond donors⢠Polymer flexibility⢠Prevents ligand exchange | Reversible cross-linker providing mechanical properties |
| Aâ-type zinc porphyrin | ⢠Ditopic coordination capability⢠Rigid planar structure⢠Tunable electronic properties | Structural linker with optoelectronic functionality |
The modular ICSs approach enables researchers to independently optimize different material functions by selecting appropriate NBBs and organic components from a growing library of well-characterized building blocks. This "mix-and-match" strategy dramatically accelerates materials development compared to the de novo synthesis required for conventional materials.
The tunability of ICSs represents perhaps their most significant advantage over conventional materials, enabling precise control over material properties through systematic variation of building blocks and their interactions. This fine-tuning capability operates at multiple levelsâfrom electronic structure to nanoscale organizationâallowing researchers to engineer materials with customized characteristics for specific applications.
The electronic properties of ICSs can be modulated through strategic selection of inorganic clusters and their organic counterparts:
Cluster Core Manipulation: The intrinsic properties of inorganic clusters, such as the photochromic behavior of titanium oxo-clusters, can be preserved while tailoring their surface chemistry for compatibility with organic matrices [111]. This decoupling of core functionality from interfacial interactions enables independent optimization of different material properties.
Ligand Field Engineering: In supramolecular coordination complexes (SCCs), the geometric and electronic properties of metal nodes can be fine-tuned by employing different metal ions (Pd(II), Pt(II), Au(III), etc.) with complementary multidentate ligands [6]. This approach allows control over the stability, solubility, and redox characteristics of the resulting assemblies.
Dynamic Behavior Control: The reversibility of supramolecular interactions introduces a temporal dimension to tunability. Systems based on classic coordination chemistry typically exhibit dynamic behavior with reversible formation and disruption, while those incorporating late transition metal N-heterocyclic carbene (NHC) complexes form more static structures due to the strong metal-carbon bonds [6].
The physical properties of ICSs can be precisely adjusted to meet application-specific requirements:
Cross-link Density Control: In hybrid supramolecular networks incorporating Tiââ-pyrrolidone clusters and telechelic PDMS, the material's mechanical state (gel vs. viscous liquid) can be tuned by varying the polymer chain length between cross-links [111]. Shorter chains (PDMS1, ~8000 g/mol) produce rigid gels, while longer chains (PDMS2, ~35000 g/mol) yield viscous liquids, demonstrating control over material mechanics through modular design.
Architectural Diversity: The geometry of ICSs can be controlled through the coordination preferences of metal nodes and the design of organic linkers. The paneling method (face-directed approach) and edge-directed approach offer complementary pathways to two-dimensional polygons and three-dimensional polyhedra with defined cavities [6]. Recent synthesis of 'PdââLââ'âthe largest discrete self-assembled edge-directed polyhedron reportedâdemonstrates the scalability of these architectures [6].
Stimuli-Responsive Behavior: The dynamic nature of supramolecular interactions introduces inherent responsiveness to external stimuli. Changes in temperature, pH, or competitive solvents can reversibly alter material properties, enabling applications in sensing, drug delivery, and adaptive materials [114].
Table 2: Tunability Parameters in ICSs
| Tuning Parameter | Experimental Control | Resulting Property Change |
|---|---|---|
| Cluster Functionality | Number of H-bond acceptors/donors | Association strength and network connectivity |
| Polymer Chain Length | Molecular weight of telechelic polymer (8K vs. 35K Da) | Mechanical state (gel vs. viscous liquid) |
| Metal Node Geometry | Coordination geometry (linear, square planar, octahedral) | Overall architecture (2D polygon, 3D cage) |
| Ligand Electronic Properties | Backbone or wingtip modification of NHC donors | Electron density at metal center, stability |
| Building Block Ratio | Stoichiometry of complementary components | Network connectivity and defect concentration |
The following diagram illustrates the workflow for tuning ICSs properties through systematic building block variation:
Diagram: ICSs Property Tuning Workflow
ICSs excel at combining multiple functionalities within a single material systemâa challenging feat for conventional materials. This multifunctionality often arises from the synergistic combination of inorganic and organic components, each contributing distinct properties that are integrated through supramolecular assembly. The resulting materials exhibit emergent properties not possessed by their individual components.
The strategic combination of inorganic clusters and organic polymers produces materials with enhanced capabilities:
Photochromism with Mechanical Integrity: Titanium oxo-cluster/PDMS hybrids demonstrate how ICSs can maintain the intrinsic photochromic properties of the inorganic component (UV-induced darkening due to Ti³⺠formation) while gaining improved mechanical characteristics from the polymer matrix [111]. This combination enables the creation of light-responsive materials with structural integrity.
Catalytic Activity with Molecular Recognition: Supramolecular coordination complexes (SCCs) with defined cavities combine catalytic metal centers with selective molecular recognition properties [6]. These systems can preferentially bind and transform specific substrates, mimicking enzymatic efficiency and selectivity.
Luminescence with Processability: Molybdenum halide clusters functionalized with pyridine groups maintain their notable photophysical properties when assembled with zinc porphyrins, creating luminescent hybrid materials with solution processability [112]. Such integration is particularly valuable for optoelectronic applications.
The reversibility of supramolecular interactions introduces adaptive capabilities not found in conventional covalently-linked materials:
Self-Healing Characteristics: The dynamic nature of hydrogen bonds in Tiââ-pyrrolidone/PDMS networks enables potential self-repair capabilities, as broken connections can reform after damage [111]. This property emerges from the collective behavior of numerous reversible interactions rather than any single molecular component.
Stimuli-Responsive Structural Changes: Systems based on host-guest chemistry, such those employing cucurbit[8]uril as a molecular "host," can undergo reversible assembly and disassembly in response to chemical, light, or thermal stimuli [114]. This adaptability enables applications in controlled release and sensing.
Adaptive Mechanical Properties: The frequency-dependent behavior of supramolecular polymer networksâwhere material response varies with applied stress rateâprovides inherent damping characteristics valuable for coatings and biomedical applications [114].
The synthesis and analysis of ICSs requires specialized methodologies that account for their hybrid nature and reliance on non-covalent interactions. The following protocols and characterization techniques have been validated through pioneering studies in the field.
Protocol 1: Functionalization of Titanium Oxo-Cluster with Hydrogen-Bond Acceptors
This procedure modifies TiââOââ(OEt)ââ clusters to introduce pyrrolidone ligands as hydrogen-bond acceptors [111]:
Protocol 2: Preparation of Thiourea-Terminated PDMS Hydrogen-Bond Donors
This synthesis produces telechelic polymers with strong hydrogen-bond donating end groups [111]:
Protocol 3: Assembly of Hybrid Supramolecular Network
This general method produces ICSs through supramolecular interactions between functionalized clusters and polymers [111]:
Table 3: Key Reagents for ICSs Research
| Reagent/Material | Function in ICSs Research | Specific Example |
|---|---|---|
| Metal Alkoxide Precursors | Source of metal ions for cluster formation | Ti(OEt)â for titanium oxo-clusters |
| Telechelic Polymers | Flexible spacers with end-group functionality | Aminopropyl-terminated PDMS (8K-35K Da) |
| Molecular Recognition Groups | Specific supramolecular interactions | 1-(2-hydroxyethyl)-2-pyrrolidone (H-bond acceptor), Phenyl isothiocyanate (for thiourea formation) |
| Solvents | Reaction medium for assembly | Anhydrous toluene, Dichloromethane |
| Spectroscopic Probes | Verification of supramolecular interactions | Deuterated chloroform for NMR, KBr for FTIR pellets |
The analysis of ICSs requires multimodal characterization to verify both structure and supramolecular interactions:
Nuclear Magnetic Resonance (NMR) Spectroscopy: ¹H NMR confirms functionalization degree and purity. ¹â·O NMR verifies inorganic core integrity after surface modification. Diffusion Ordered Spectroscopy (DOSY) assesses molecular assembly in solution [111] [112].
Fourier-Transform Infrared (FTIR) Spectroscopy: Identifies hydrogen bonding through shifts in N-H stretching frequencies (free N-H: 3410-3393 cmâ»Â¹; H-bonded N-H: 3250-3300 cmâ»Â¹) [111].
Single-Crystal X-ray Diffraction (SCXRD): Determines atomic-level structure of crystalline ICSs, as demonstrated for molybdenum cluster-zinc porphyrin hybrids [112].
Rheological Analysis: Quantifies mechanical properties and gelation behavior of supramolecular networks [111].
The following diagram illustrates the integrated workflow for ICSs synthesis and characterization:
Diagram: ICSs Synthesis and Characterization Workflow
The advantages of ICSs become most apparent when directly compared to traditional material systems. The following analysis highlights the transformative potential of the supramolecular approach across critical material parameters.
Table 4: Performance Comparison: ICSs vs. Conventional Materials
| Material Property | ICSs Performance | Conventional Materials Performance | Key Advantage |
|---|---|---|---|
| Structure-Property Relationship | Precise control through building block selection | Limited to compositional variations | Molecular-level design capability |
| Processability | Soluble precursors enable solution processing | Often require harsh processing conditions | Mild assembly conditions |
| Functional Integration | Innate multifunctionality from hybrid composition | Typically require composite strategies | Synergistic property combinations |
| Stimuli Responsiveness | Intrinsic through reversible bonds | Limited to specially designed polymers | Adaptive and dynamic behavior |
| Self-Healing Potential | Built-in through supramolecular dynamics | Rare and often requires external triggers | Autonomous repair capability |
| Structural Precision | Defined molecular architectures | Mostly amorphous or semi-crystalline | Nanoscale order and predictability |
The comparative analysis reveals that ICSs outperform conventional materials across multiple dimensions, particularly when application requirements include multifunctionality, adaptability, or precise structure-property relationships. The modular design paradigm enables researchers to independently address different material functions through appropriate building block selection, avoiding the performance trade-offs common in traditional material systems.
Inorganic Cluster Supramolecular compounds represent a paradigm shift in materials design, leveraging modularity, tunability, and multifunctionality to overcome limitations of conventional material systems. The supramolecular approachâemploying specific, reversible interactions to direct the assembly of well-defined inorganic and organic componentsâenables unprecedented control over material architecture and properties. As demonstrated by titanium oxo-cluster/PDMS hybrids and molybdenum cluster/porphyrin systems, ICSs can integrate complementary characteristics from different material classes, producing synergistic combinations of mechanical, optical, electronic, and responsive properties.
The future development of ICSs will likely focus on several promising directions. First, expanding the library of inorganic NBBs to include clusters with diverse electronic, magnetic, and catalytic properties will broaden the functional scope of these materials. Second, incorporating more sophisticated supramolecular motifsâsuch as mechanically interlocked architectures or biologically-inspired recognition elementsâwill enhance the complexity and precision of ICSs assembly. Third, advancing theoretical models to predict the structure-property relationships of ICSs will accelerate the rational design of materials with tailored characteristics.
For researchers in drug development and biomedical applications, ICSs offer particularly compelling opportunities. The host-guest chemistry of metallacages can be exploited for drug delivery, while the dynamic nature of supramolecular interactions enables stimuli-responsive release mechanisms [6]. The integration of imaging functionalities (luminescence, MRI contrast) with therapeutic capabilities in single ICSs platforms aligns perfectly with the emerging field of theranostics [6]. As fundamental understanding of supramolecular assembly principles deepens and synthetic methodologies advance, ICSs are poised to become indispensable tools for creating next-generation functional materials that transcend the limitations of conventional systems.
The exploration of inorganic cluster supramolecular compounds marks a paradigm shift in materials science and biomedical research. Synthesizing knowledge across all intents confirms that the rational design and precise synthesis of ICSs, guided by fundamental coordination principles, enable unprecedented control over their physical and chemical properties. Their demonstrated success in targeted drug delivery, catalysis, and theranostics underscores a significant advantage over many conventional materials due to their inherent modularity and multifunctionality. Future directions must focus on deepening the understanding of their in vivo behavior and long-term biocompatibility, optimizing large-scale synthesis for clinical translation, and exploring innovative applications in diagnostics and combination therapies. The continuous refinement of synthetic and analytical methodologies will undoubtedly unlock the full potential of ICSs, solidifying their role as next-generation smart materials in advanced therapeutics and beyond.