Fact-checked by Grok 2 weeks ago

Stability constants of complexes

Stability constants of coordination complexes, also known as formation constants, are constants that quantify the propensity of a (or central ) and to form a stable in , expressing the strength of their association. These constants can be stepwise, describing the addition of each successive (denoted as K_n), or cumulative (overall), representing the complete formation of the (denoted as \beta_n), and larger values indicate more stable species relative to their dissociated components. For a general mononuclear formation \ce{M + nL ⇌ ML_n}, the overall stability constant is given by \beta_n = \frac{[\ce{ML_n}]}{[\ce{M}][\ce{L}]^n}, where M is the and L is the . The determination of stability constants is essential for understanding complex behavior in aqueous environments, influencing applications in , , pharmaceuticals, and biological systems where metal-ligand interactions govern reactivity and selectivity. Methods for measuring these constants include potentiometry, , and techniques, with values often reported in logarithmic form (e.g., \log \beta) for across wide ranges. Notable examples include the highly stable silver-cyanide [\ce{Ag(CN)2}]^- with \log K_f \approx 21 and the copper-ammonia [\ce{Cu(NH3)4}]^{2+} with stepwise constants decreasing from \log K_1 \approx 4.0 to \log K_4 \approx 2.1, illustrating how ligand binding affinity diminishes with . Factors affecting stability include the nature of the metal (e.g., higher charge and smaller enhance stability), ligand (chelate effect from multidentate ligands like EDTA increases \beta by 10^4–10^14 compared to monodentate analogs due to gains), and solution conditions such as and . Thermodynamic stability reflects positioning, while kinetic stability pertains to reaction rates, both critical for predicting complex durability in practical scenarios like metal ion extraction or .

Fundamentals

Definition and significance

Stability constants, also known as formation constants and denoted as β or K, are equilibrium constants that measure the extent of complex formation between a metal ion and one or more ligands in solution, indicating the strength of the metal-ligand bond under given conditions. For the prototypical reaction where a metal ion M^{n+} reacts with n equivalents of a ligand L^{m-} to form the complex ML_n, the overall stability constant β_n is defined as β_n = \frac{[ML_n]}{[M][L]^n}, where concentrations are typically expressed at equilibrium. These constants provide a quantitative basis for assessing how readily a complex forms, with larger values signifying greater stability and resistance to dissociation. The significance of stability constants lies in their ability to predict the , , and reactivity of metal ions in aqueous environments, enabling chemists to model the distribution of species in complex mixtures. In coordination chemistry, they elucidate binding preferences and effects, guiding the design of synthetic complexes for and materials. Within , stability constants underpin complexometric titrations, such as those using EDTA for precise metal ion quantification, where high constants ensure selective and complete complexation. In , they are vital for simulating metal pollutant transport, , and remediation, as seen in modeling the of like mercury in natural waters. The foundational measurements of stability constants were established in the 1940s by Jannik Bjerrum, who developed methods to determine them for metal-ammine complexes in aqueous solution. Comprehensive databases, such as the NIST Critically Selected Stability Constants, compile these values to support ongoing research across disciplines.

Historical development

The recognition of complex formation dates back to early observations in the 18th and 19th centuries, where qualitative evidence of coordination compounds emerged through pigment synthesis and solution chemistry experiments. Prussian blue, discovered in 1704 by Johann Jacob Diesbach as a deep blue pigment from the reaction of iron salts with potassium ferrocyanide, represented one of the earliest known coordination compounds, formulated as Fe₄[Fe(CN)₆]₃·xH₂O, though its crystal structure was first elucidated in 1936, with further refinement in 1977. In the 19th century, precursors to Alfred Werner's coordination theory noted distinct behaviors in solution, such as Tassaert's 1798 observation of brown solutions from cobalt(II) salts and ammonia upon air exposure, indicating ammine complex formation. Further studies by Genth in 1851 and Frémy in 1852 on cobalt ammine complexes, including isolation of [Co(NH₃)₅Cl]Cl₂ and color-based nomenclature for species like luteocobalt, highlighted stepwise ligand binding without quantitative equilibrium analysis. Quantitative approaches to stability constants began in the early , with pioneering work in the –1940s focusing on stepwise formation in . Niels Bjerrum and Jaques independently proposed principles for stepwise equilibrium constants in 1914, but Jannik Bjerrum advanced this in the 1940s through experimental and theoretical studies on metal-ammine systems. His 1941 doctoral thesis and book, Metal Ammine Formation in Aqueous Solution, introduced systematic potentiometric methods using glass electrodes to determine stepwise stability constants (log K₁ to log K₆) for complexes like [Cu(NH₃)ₙ]²⁺, resolving computational challenges via formation functions (n-bar plots). I. Leden complemented this in the 1940s with polarographic techniques for labile systems, enabling broader application to transition metal ions. The 1950s–1970s marked a shift to refined experimental methods and , driven by potentiometric titrations and international collaboration. Gerold Schwarzenbach's 1952 work recognized the chelate effect, demonstrating enhanced stability of multidentate ligands like over monodentate due to gains, as seen in log K values for [Ni(en)₃]²⁺ exceeding those for [Ni(NH₃)₆]²⁺. Irving and Rossotti's 1953 method simplified calculations of successive stability constants from pH titration curves, widely adopted for proton-ligand and metal-ligand . The International Union of Pure and Applied Chemistry (IUPAC) initiated in 1957 with Stability Constants of Metal-Ion Complexes (Special Publications No. 6 and 7), edited by Bjerrum, Schwarzenbach, and Sillén, compiling critically evaluated for over 1,000 systems. Potentiometric methods proliferated in the 1960s–1970s, supported by early computer programs like LETAGROP (1964) for error minimization. Key milestones included the 1970s launch of IUPAC's on Equilibrium Data for ongoing compilations and the inception of formation constants databases, facilitating global data access and modeling.

Theoretical Framework

Stepwise and overall stability constants

In the formation of metal- complexes involving multiple ligands, the process often occurs stepwise, where each successive binds to the partially coordinated metal center. The stepwise stability constant, denoted as K_i, quantifies the for the addition of the i-th to the complex already containing i-1 ligands. For a general monodentate L and metal M, this is expressed as: \mathrm{M}L_{i-1} + \mathrm{L} \rightleftharpoons \mathrm{M}L_i \quad K_i = \frac{[\mathrm{M}L_i]}{[\mathrm{M}L_{i-1}][\mathrm{L}]} These constants reflect the incremental stability gained at each coordination step. The overall stability constant, also known as the cumulative or formation constant \beta_n, describes the equilibrium for the complete formation of the complex with n ligands from the free metal ion and ligands. It is defined as: \mathrm{M} + n\mathrm{L} \rightleftharpoons \mathrm{M}L_n \quad \beta_n = \frac{[\mathrm{M}L_n]}{[\mathrm{M}][\mathrm{L}]^n} The overall constant is the product of the individual stepwise constants: \beta_n = K_1 K_2 \cdots K_n. This relationship allows the overall stability to be derived from stepwise data or vice versa. Due to the large magnitudes of stability constants, they are commonly reported in logarithmic form: \log K_i for stepwise constants and \log \beta_n for the overall constant. The logarithmic relationship follows as \log \beta_n = \log K_1 + \log K_2 + \cdots + \log K_n, facilitating easier comparison and calculation in experimental analyses. A typical pattern observed in stepwise constants for monodentate ligands is K_1 > K_2 > \cdots > K_n, with each subsequent constant decreasing. This trend arises partly from statistical factors: as more ligands bind, the number of available coordination sites decreases while the number of bound ligands that must be displaced or compete increases. For an octahedral complex, statistical predicts approximate ratios such as K_i / K_{i+1} \approx (n - i + 1)/i, contributing to the observed diminution, though electronic and also play roles. A representative example is the formation of the tetraamminecopper(II) complex, [Cu(NH₃)₄]²⁺, in at 25°C. The stepwise constants are log K_1 = 4.3, log K_2 = 3.6, log K_3 = 3.0, and log K_4 = 2.3, yielding an overall log \beta_4 = 13.2. These values illustrate the decreasing stability with successive coordination, consistent with both statistical expectations and experimental potentiometric data.

Types of complexes

In , metal s often form products, which are a key type of complex characterized by stability constants. These arise from reactions where coordinated molecules deprotonate, yielding such as \mathrm{M(OH)_p^{n-p}} for a metal \mathrm{M^{n+}}. The overall stability constant for such is defined as \beta_p = \frac{[\mathrm{M(OH)_p^{n-p}}][\mathrm{H^+}]^p}{[\mathrm{M^{n+}}]}, where p indicates the number of ligands incorporated. This formulation accounts for the involvement of protons in the , distinguishing complexes from simple \mathrm{ML_n} where L is a non- ; here, acts implicitly as the source of both the ligand (OH^-) and the competing H^+. Stepwise stability constants can describe multi-step processes, such as sequential deprotonations leading to polymeric or higher-order . A representative example is the of \mathrm{Al^{3+}} in , which forms mononuclear like \mathrm{Al(OH)^{2+}} and \mathrm{Al(OH)_2^{+}}, alongside polynuclear forms such as \mathrm{Al_2(OH)_2^{4+}}. At 25°C and zero , the overall constants include \log \beta_1 = -4.98 for \mathrm{Al^{3+}} + \mathrm{H_2O} \rightleftharpoons \mathrm{Al(OH)^{2+}} + \mathrm{H^+}, \log \beta_2 = -10.63 for the dihydroxy , and \log \beta_{2,2} = -7.62 for the dimer, reflecting increasing stability with successive hydroxo coordination but also the tendency toward at higher . These values highlight how hydrolysis stability decreases with proton concentration, influencing aluminum in natural waters and industrial processes. Another category encompasses acid-base complexes, where ligands can , affecting their coordination to metals. The protonation constant for a is given by K_H = \frac{[\mathrm{HL}]}{[\mathrm{H^+}][\mathrm{L^-}]}, measuring the ligand's basicity and its competition with metal binding. Mixed proton-metal species, such as \mathrm{MHL}, further complicate equilibria, with stability constants incorporating both and coordination steps; for instance, the constant for \mathrm{ML} + \mathrm{H^+} \rightleftharpoons \mathrm{MHL} is \beta_{\mathrm{MHL}} = \frac{[\mathrm{MHL}]}{[\mathrm{ML}][\mathrm{H^+}]}. Unlike pure , these involve explicit ligand protonation, often seen in systems with polyprotic ligands like carbonates or carboxylates. Carbonate complexes exemplify this type, forming species like \mathrm{MCO_3^{+}} for trivalent metals, where the ligand's protonation to \mathrm{HCO_3^-} modulates stability. The mixed proton form is expressed as \mathrm{CO_3H}\beta_1 = \frac{[\mathrm{MCO_3^{+}}][\mathrm{H^+}][\mathrm{M^{3+}}]^{-1}[\mathrm{HCO_3^-}]^{-1}, with \log \mathrm{CO_3H}\beta_1^0 \approx 2.85 for \mathrm{Y^{3+}} at 25°C and zero ionic strength, indicating moderate stability influenced by the carbonate's acid-base behavior. Similarly, for \mathrm{La^{3+}}, \log \mathrm{CO_3H}\beta_1^0 = 3.60, showing lanthanide contraction effects on binding strength. These complexes are crucial in geochemical contexts, where proton activity alters speciation.

Thermodynamic Considerations

Ionic strength dependence

The stability of metal-ligand complexes in aqueous solutions is influenced by the (I) of the medium, which affects the apparent stability constants (K') measured in terms of concentrations. These apparent constants differ from the thermodynamic stability constants (β), which are defined in terms of activities at infinite dilution (I = 0) and are independent of ionic strength. The relationship arises from non-ideal behavior in solutions, where ionic interactions alter the effective concentrations through activity coefficients (γ), such that β = K' × (γ_M γ_L^n / γ_{ML_n}), with subscripts denoting the metal ion (M), (L), and (ML_n). Activity coefficients account for electrostatic interactions among ions, and their ionic strength dependence is modeled using theories derived from Debye-Hückel theory. In the limiting case for dilute solutions (I < 0.01 mol·kg⁻¹), the extended Debye-Hückel equation approximates log γ = -A z^2 √I / (1 + √I), where A is the Debye-Hückel constant (≈0.51 at 25°C in water), z is the ion charge, and I is the on the molal scale. This model captures long-range electrostatic effects but underestimates deviations at higher I due to short-range interactions. For a broader range of moderate ionic strengths (up to I ≈ 0.1–0.2 mol·kg⁻¹), the empirical Davies equation provides a better fit: log γ = -A z^2 [√I / (1 + √I) - 0.3 I]. This extension incorporates a linear term to account for additional salting-out effects, improving accuracy for 1:1 electrolytes by about ±0.1 log units in activity coefficients compared to the basic Debye-Hückel form. At higher ionic strengths (I > 0.5 mol·kg⁻¹), such as in seawater or concentrated brines, the Specific Ion Interaction Theory (SIT) is preferred, as it separates long-range Debye-Hückel effects from short-range specific ion pairing: log γ_j = -z_j^2 D + ∑ ε(j, k) m_k, where D = A √I / (1 + 1.5 √I), ε(j, k) are empirical interaction coefficients, and m_k are molalities of background ions. For complex formation equilibria, the apparent log K' relates to the thermodynamic log β by log β = log K' + Δz^2 D - Δε I, with Δz^2 = z_{ML_n}^2 - z_M^2 - n z_L^2 and Δε incorporating the relevant ε terms. This approach enables reliable extrapolation to I = 0 even from data at high I. A representative example is the Cu^{2+}-EDTA complex (Cu(EDTA)^{2-}), where the overall stability constant (log β) decreases with increasing I due to the high charges involved (Δz^2 = -16). Using SIT with NaCl media, the interaction parameter ε(Na^+, EDTA^{4-}) ≈ 0.32 kg·mol^{-1} leads to a shift of approximately -2 to -3 log units in log β from I = 0 to I = 3 mol·kg^{-1}, highlighting the need for corrections in applications like or environmental speciation.

Temperature dependence

The temperature dependence of stability constants for metal complexes arises from the thermodynamic parameters governing the complexation equilibrium, primarily the standard enthalpy change (ΔH°) and change (ΔS°). According to the van't Hoff equation, the variation of the logarithm of the stability constant β with temperature T is given by \frac{d(\ln \beta)}{dT} = \frac{\Delta H^\circ}{RT^2}, where R is the . This relationship indicates that the temperature sensitivity of β depends directly on the sign and magnitude of ΔH° for the complex formation reaction. Assuming ΔH° is constant over a limited temperature range, the equation can be integrated to yield \ln\left(\frac{\beta_{T_2}}{\beta_{T_1}}\right) = -\frac{\Delta H^\circ}{R} \left( \frac{1}{T_2} - \frac{1}{T_1} \right). This form allows prediction of stability constants at different temperatures if β and ΔH° are known at one temperature, facilitating corrections in modeling. The sign of ΔH° determines whether stability increases or decreases with rising temperature. For hard acid-hard base pairs, such as those following the hard-soft acid-base (HSAB) principle (e.g., Mg^{2+} with oxygen donors), complex formation is typically exothermic (ΔH° < 0), resulting in a decrease in β as temperature increases because the forward reaction is enthalpically favored at lower temperatures. In contrast, soft acid-soft base interactions, like Hg^{2+} with iodide, often exhibit endothermic or near-zero ΔH° values, leading to stability that either increases with temperature or shows minimal variation, as the reaction benefits less from enthalpic contributions. Typical ΔH° magnitudes range from -20 to -50 kJ/mol for exothermic hard-hard complexes and +5 to +20 kJ/mol for endothermic soft-soft ones, influencing applications in processes sensitive to thermal conditions. Entropy contributions (ΔS°) play a crucial role in the overall temperature dependence, particularly through changes in solvation. Complex formation often involves the release of coordinated water molecules from the metal ion's hydration sphere and ligand solvation shell, yielding positive ΔS° values (typically 20–100 J mol^{-1} K^{-1}) that enhance stability at higher temperatures via the -TΔS° term in the . For example, in nickel(II)-ethylenediamine (Ni^{2+}-en) complexes, the stepwise formation shows favorable entropy gains (ΔS° ≈ 30–50 J mol^{-1} K^{-1} per step) due to desolvation, offsetting sometimes modest enthalpic changes and contributing to the observed stability even as temperature rises. In practice, stability constants are standardized at 25°C (298.15 K) to ensure comparability across datasets, as recommended in critical compilations. For applications requiring data at other temperatures, such as environmental modeling or industrial processes in the 0–50°C range, corrections are applied using measured or estimated ΔH° values, with the van't Hoff integration providing reliable extrapolations within this narrow interval where ΔH° approximations hold.

Factors Affecting Stability

Chelate and macrocyclic effects

The chelate effect describes the increased thermodynamic stability of metal complexes formed with multidentate ligands relative to analogous complexes with separate monodentate ligands, driven primarily by a favorable change (ΔS > 0) resulting from the liberation of fewer independent molecules upon . This phenomenon, first systematically explored by Schwarzenbach in 1952, arises because chelate formation releases a smaller number of solvent molecules (typically ) compared to stepwise binding of individual s, reducing translational entropy loss in the system. For instance, the hexadentate ligand ethylenediaminetetraacetate (EDTA) forms a highly stable complex with Cu^{2+}, with an overall stability constant of log β ≈ 18.8 (at 25°C, I = 0.1 M), far exceeding the cumulative stepwise constant for [Cu(NH_3)_4]^{2+} (log β_4 ≈ 12.6), as the single EDTA molecule displaces six molecules from the aquo ion whereas four NH_3 ligands would displace eight. A key quantitative measure of the chelate effect is the enhancement per ring formed, often 3–4 log units in stability constants for five- or six-membered chelate rings. This is evident in the binding of Cu^{2+} to ethylenediamine (en), where log K for [Cu(en)(H_2O)_4]^{2+} is approximately 10.6, compared to log β_2 ≈ 7.9 for [Cu(NH_3)_2(H_2O)_4]^{2+}, yielding a difference of ~2.7 units for one chelate ring; for the bis complex [Cu(en)_2]^{2+}, the overall log β_2 reaches ~20.0 versus ~12.6 for [Cu(NH_3)_4]^{2+}, confirming ~3.7 units per ring. Similar enhancements occur in other first-row transition metal systems, such as Ni^{2+}, where log β_6 for [Ni(en)_3]^{2+} is 18.3 compared to 8.7 for [Ni(NH_3)_6]^{2+}, highlighting the entropic advantage of ring closure over multiple independent ligand exchanges. The macrocyclic effect builds upon the chelate effect by further stabilizing complexes through the preorganized, rigid cyclic structure of the ligand, which reduces the entropic penalty associated with conformational changes during binding and often confers additional kinetic inertness to the complex. Seminal work by Pedersen in 1967 on crown ethers demonstrated this for alkali metal ions, where 18-crown-6 binds K^+ with log K ≈ 6.0–6.1 (in water/ mixtures), 10^2 to 10^4 times more stable than comparable acyclic polyethers like , due to the preformed cavity matching the ion radius and minimizing desolvation costs. In azamacrocycles, cyclam (1,4,8,11-tetraazacyclotetradecane) exemplifies the effect with Cu^{2+}, yielding log β ≈ 25.3–25.9 for [Cu(cyclam)]^{2+}, versus log β ≈ 20.4 for the acyclic tetradentate analog (N(CH_2CH_2NH_2)_3), a stability increase of ~10^5 attributed to the locked conformation and reduced ligand flexibility. Representative examples include smaller-ring macrocycles like aneN_3 (1,4,7-triazacyclononane), where the Cu^{2+} exhibits log K ≈ 15.5 and elevated proton-transfer barriers (>13 kcal/mol) compared to the acyclic tridentate dien (diethylenetriamine, log K ≈ 12.5), illustrating the macrocyclic effect's role in enhancing both thermodynamic and kinetic stability through enforced planarity and reduced vibrational freedom. This kinetic inertness, beyond mere thermodynamic gain, is a hallmark of macrocycles, slowing dissociation rates by orders of magnitude relative to open-chain analogs. Crown ethers and cyclams thus provide models for entropy-driven preorganization, with stability enhancements scaling with ring size match to the metal ion. In polyaminocarboxylates such as EDTA, is further modulated by pH-dependent of the groups (–COOH), which at high (>10) fully converts the to its Y^{4-} form, maximizing negative and donor site availability for metal coordination. This enhances conditional constants by 10^3–10^6 relative to protonated forms at lower , as seen in Cu^{2+}-EDTA where effective binding requires the tetraanionic to compete with ; without it, equilibria reduce the free concentration. Geometric constraints in multidentate ligands can amplify these entropy-driven effects by enforcing optimal donor orientations.

Geometrical and metal ion factors

The of coordination complexes is profoundly influenced by geometrical factors, which arise from the compatibility between the preferred of the metal ion and the spatial arrangement of ligand donor atoms. In complexes, the crystal field stabilization (CFSE) provides a quantitative measure of this influence, as the splitting of d-orbitals in specific geometries lowers the overall electron configurations. For instance, d³ metals like Cr³⁺ exhibit particularly high CFSE (-1.2 Δ_o) in octahedral environments, favoring six-coordinate structures and enhancing complex compared to other geometries. The size of the metal ion, quantified by its ionic radius, also governs stability through electrostatic interactions and bond strength. Smaller ions possess higher charge density, leading to shorter metal-ligand bonds and greater stability for a given ligand and charge. Fe³⁺, with an ionic radius of 0.65 Å (high-spin, coordination number 6), forms significantly more stable complexes than the larger Cd²⁺ (0.95 Å, coordination number 6), as the higher charge-to-radius ratio of Fe³⁺ strengthens ionic and covalent contributions to bonding. This effect is exemplified in their respective EDTA complexes, where the overall stability constant reflects the superior binding of the smaller, higher-charge-density ion. The Hard-Soft Acid-Base (HSAB) principle elucidates how the polarizability and charge distribution of metal ions and ligands dictate stability, with hard-hard and soft-soft pairings yielding more stable complexes than hard-soft mismatches. Hard acids, such as Al³⁺ (small size, high charge density), preferentially bind hard bases like F⁻, forming robust complexes due to favorable electrostatic interactions. Conversely, soft acids like Hg²⁺ (large, polarizable) form exceptionally stable bonds with soft bases such as I⁻, where the overall stability constant for [HgI₄]²⁻ reaches log β₄ ≈ 30, far exceeding values for mismatched pairs. For divalent ions of the first-row transition metals, the Irving-Williams series captures a systematic stability trend: Mn²⁺ < Fe²⁺ < Co²⁺ < Ni²⁺ < Cu²⁺ > Zn²⁺. This order stems from the interplay of decreasing ionic radii (enhancing ) and increasing CFSE across the series, with a maximum at Cu²⁺ due to its d⁹ configuration and partial filling of the e_g orbitals. The series is prominently observed in complexes with multidentate ligands like EDTA, where stability constants follow the same progression, underscoring the role of d-electron effects in metal ion selectivity.
Metal IonIonic Radius (Å, CN=6)log β ([M(EDTA)]^{2-}, 25°C, I=0.1 M)
Mn²⁺0.8313.9
Fe²⁺0.7814.3
Co²⁺0.74516.3
Ni²⁺0.6918.6
Cu²⁺0.7318.8
Zn²⁺0.7416.7
These log β values, critically evaluated from potentiometric and spectroscopic data, illustrate the Irving-Williams trend quantitatively, with Cu²⁺ exhibiting the highest due to optimal geometrical and factors.

Experimental Methods

Classical determination techniques

Classical determination techniques encompass a range of empirical methods developed in the mid-20th century to measure constants of metal-ligand complexes in , relying on observable changes in physical properties during or establishment. These approaches, pioneered in the , provide direct experimental on formation equilibria without computational modeling, though they require careful control of variables like and temperature to ensure accuracy. Potentiometric titration is one of the most widely used classical methods, involving the monitoring of pH changes as a ligand is added to a metal ion solution or vice versa, using a glass electrode to track proton release or uptake during complex formation. The stability constant \beta_{pq} for a general complex \ce{M_p L_q H_r} is derived from pH data via mass and charge balance equations, often refined through nonlinear least-squares fitting to minimize residuals between observed and calculated titration curves. Gran plots enhance this technique by linearizing data before and after equivalence points, enabling endpoint determination without relying on inflection points, which is particularly useful for weak complexes where pH changes are subtle; for instance, the plot of $10^{pH} V (where V is titrant volume) versus volume yields a straight line intersecting at the endpoint. This method has been applied extensively to EDTA-metal titrations, where the high stability of EDTA chelates (e.g., \log \beta \approx 18.8 for [\ce{Cu(EDTA)}]^{2-} at 25°C and I=0.1) allows precise determination of metal concentrations and constants, though side reactions like hydrolysis can introduce errors if not accounted for by including auxiliary ligands. Corrections for ionic strength (via activity coefficients) and temperature (typically maintained at 25±0.05°C) are essential, as variations can shift apparent constants by up to 10% per 0.1 M change in I. Spectrophotometric methods exploit shifts in UV-Vis absorption spectra upon metal-ligand binding, quantifying complex formation through changes in at specific wavelengths. In Job's method (continuous variation), equimolar solutions of metal and are mixed in varying ratios while keeping total concentration constant, and the maximum at the stoichiometric ratio (e.g., 1:1 for ML) indicates the complex geometry; stability constants are then calculated from the data assuming Beer's law holds. This approach is effective for colored complexes like those of d-transition metals with organic s, such as the \ce{Cu(II)-[histidine](/page/Histidine)} system where \log K \approx 10.5 at physiological , but it assumes no higher-order species and can be sensitive to overlapping spectra from free components. Mole-ratio methods complement this by plotting against excess, yielding both and K from the intersection of linear segments. Errors arise from non-ideal behavior at high concentrations or if side equilibria (e.g., ) alter the spectrum, necessitating spectral . Distribution or solvent extraction techniques determine stability constants by measuring the partition of metal ions or complexes between an aqueous phase and an immiscible organic solvent, using the distribution coefficient D = \frac{[\ce{M}]_\text{org}}{[\ce{M}]_\text{aq}} to derive K from equilibria involving extractable species. A ligand is added to favor complex formation, shifting the metal into one phase, and constants are calculated from extraction isotherms at varying ligand concentrations; for example, in the extraction of \ce{UO2^{2+}} with thenoyltrifluoroacetone, \log \beta_2 \approx 8.5 was obtained by fitting data to stepwise formation models. This method suits hydrophobic ligands and labile complexes but requires knowledge of extraction constants and can suffer from incomplete phase separation or emulsion formation as errors. Nuclear magnetic resonance (NMR) and (EPR) spectroscopies probe dynamic equilibria in solution by analyzing chemical shifts or relaxation rates influenced by complexation. In NMR, ligand proton signals shift upon binding due to the paramagnetic or diamagnetic effects of the metal, allowing K calculation from weighted averages of free and bound spectra; for instance, ^{1}\ce{H} NMR on \ce{VO^{2+}-glycine} yielded \log K_1 = 8.8 by integrating peak intensities at varying metal-ligand ratios. EPR is particularly useful for paramagnetic metals like Cu(II) or Mn(II), where hyperfine splitting patterns reveal and exchange rates, enabling stability assessment in systems like \ce{Cu(II)-EDTA} with \log K \approx 18.8. These techniques excel for kinetic studies but demand high-resolution spectrometers and are limited to systems without rapid exchange broadening signals; impurities causing line broadening represent a common error source. Calorimetric measurements directly quantify the enthalpy change \Delta H of complex formation by monitoring heat evolved or absorbed during titration in an isothermal calorimeter, from which stability constants are derived when combined with \Delta G = -RT \ln K to obtain \Delta S. Direct titration calorimetry involves injecting ligand into a metal solution and fitting the heat peaks to a binding isotherm, as applied to \ce{Ca^{2+}-citrate} where \Delta H = -5.2 kJ/mol supported the measured \log \beta = 4.8. This method provides thermodynamic profiles but requires precise heat capacity calibration and is prone to baseline drift errors from side reactions like dilution heats.

Computational prediction approaches

Computational approaches to predict stability constants of metal-ligand complexes have advanced significantly, enabling the estimation of thermodynamic parameters without direct experimentation. These methods leverage , , and data-driven models to compute the of complexation, ΔG, which relates to the overall stability constant β via the equation ΔG = -RT ln β, where R is the and T is . Such predictions are particularly valuable for exploring systems where experimental data is scarce or challenging to obtain. Quantum chemical methods, primarily (DFT), provide a foundational tool for calculating the electronic structure and binding energies of complexes. The B3LYP functional, combined with basis sets like LANL2DZ or def2-TZVP, is widely used to compute the complexation energy ΔE, which approximates ΔG when corrected for and effects. is typically modeled using polarizable continuum models (PCM) or conductor-like screening models () to account for solvent interactions, improving predictions in aqueous environments. For instance, DFT calculations on metal-ammonia complexes have yielded relative constants with accuracies of 2-4 kcal/mol (corresponding to ~1-2 log units in β), demonstrating reliability for hard-hard interactions but highlighting challenges in capturing dispersion forces for softer ligands. Molecular dynamics (MD) simulations extend these predictions to dynamic solution-phase behavior, often employing (FEP) to quantify binding affinities. In FEP-MD, the difference between unbound and bound states is computed by gradually transforming the system, yielding ΔG values directly convertible to log β. This approach has been applied to macrocyclic complexes, such as [2.2.2]-cryptands with and alkaline metals, achieving mean absolute errors of ~3 kcal/mol against experimental stability constants when combined with explicit and force fields like OPLS. FEP excels in modeling conformational flexibility and solvent effects but requires substantial computational resources for larger systems. Machine learning (ML) models, particularly neural networks, offer rapid predictions by training on curated databases of experimental log β values. Graph neural networks or message-passing architectures, such as those in Chemprop, process structures via SMILES notation alongside metal descriptors to forecast stability constants, with post-2020 advancements achieving root-mean-square errors (RMSE) of 0.6-0.8 log units on diverse datasets. For example, the LOGKPREDICT model has been used to predict lanthanide-aminocarboxylate constants, such as those for EDTA complexes across La³⁺ to Lu³⁺, identifying selectivity trends with RMSE ~0.25 for relative values and accuracies within 0.5 log units for simple 1:1 systems. These models validate well against classical experimental data, often reproducing trends within 1 log unit. The primary advantages of these computational methods lie in their ability to probe inaccessible systems, such as complexes with radioactive metals like actinides (e.g., Ac³⁺ with NOTA derivatives via DFT), where experimental handling poses risks, and to address data gaps for soft s like thiols, where can interpolate from limited training sets to guide ligand design. However, challenges persist in achieving sub-0.5 log unit precision for polydentate or soft-soft interactions due to limitations or training data biases.

Data Resources

Critically evaluated data collections

Critically evaluated data collections provide vetted compilations of stability constants, ensuring reliability through rigorous and consistency checks across experimental datasets. The International Union of Pure and Applied Chemistry (IUPAC), via its Subcommittee on Solubility and Equilibrium Data, has produced peer-reviewed critical surveys since the 1970s, evaluating published equilibrium data for metal complexes to assess methodological consistency and recommend values with associated uncertainties. These surveys prioritize high-quality measurements, often reconciling discrepancies from techniques like potentiometry and , and have resulted in technical reports covering specific ligand classes, such as ethers and complexones. Complementing IUPAC efforts, the National Institute of Standards and Technology (NIST) maintains the Critically Selected Stability Constants of Metal Complexes database (SRD 46, version 8.0), which includes over 112,000 entries for more than 6,000 ligands, with constants flagged for reliability based on comprehensive literature reviews up to 2003. Although the official distribution was discontinued in , version 8.0 remains widely accessible via open-source interfaces and web tools, such as the Stability Constant Explorer, facilitating rapid bibliography searches and integration into environmental models. This resource emphasizes thermodynamic consistency and excludes unreliable data, serving as a for aqueous metal-ligand interactions. Evaluation criteria in both IUPAC and NIST compilations focus on statistical agreement among independent studies, typically requiring values to align within 0.05 log units (standard deviation ≤0.05) for "recommended" status, while tentative values allow 0.05-0.2 log units; outliers are excluded if they deviate significantly due to experimental artifacts or inconsistent conditions. For example, the recommended stepwise stability constant \log K_1 for the ^{2+}- (1:1) complex is 8.6 at zero and 25°C, derived from converged potentiometric data across multiple sources. Ongoing revisions to these collections address emerging needs. These enhancements ensure the data remain applicable to complex environmental matrices while maintaining the high standards of critical selection.

Databases and software tools

Several prominent databases serve as repositories for stability constants of metal- complexes, enabling researchers to access empirical data for modeling and analysis. The IUPAC Stability Constants Database (SC-Database), developed under the auspices of the International Union of Pure and Applied Chemistry, compiles significant published metal- stability constants along with associated thermodynamic data, encompassing approximately 24,000 literature references for about 9,800 dating from 1887 to the present. This resource supports searches filtered by metal ion, structure, (I), and (T), and allows data export in formats such as CSV or text for further processing. The NIST Critically Selected Stability Constants of Metal Complexes Database (SRD 46), maintained by the National Institute of Standards and Technology, provides a curated collection of evaluated constants for metal-ligand interactions in aqueous media, emphasizing reliability for calculations. Like the IUPAC database, it enables queries by metal, , I, and T, with options for exporting datasets compatible with . Software tools complement these databases by enabling simulation, refinement, and prediction of complex . The Hyperquad suite, including Hyperquad2013 for deriving stability constants from potentiometric or spectrophotometric and HySS for simulations, allows users to model curves and distributions under varying conditions such as , I, and T. These programs process input from databases like NIST SRD 46 and support export of diagrams in graphical or tabular formats. Visual MINTEQ, an open-source geochemical model, specializes in environmental applications, computing metal , , and using built-in stability constants derived from critically evaluated sources. ECCLES (Equilibrium Calculations in Complex Ligand Systems) is a specialized tool for refining constants from potentiometric data in multicomponent systems, particularly biological fluids like , where it models competitive binding and . Online resources, such as ACD/Labs' Percepta platform, offer structure-based predictions of related properties like acid dissociation constants () that inform assessments, though direct metal- predictions require integration with empirical databases. European Union-funded initiatives under , including the EURAD project, have developed thermodynamic databases for metal in nuclear waste contexts, incorporating post-2020 updates to constants for actinides and fission products. Emerging integrations with tools enhance these resources; for instance, datasets from IUPAC and NIST have been used to train models for predicting overall stability constants (log β), achieving mean absolute errors below 1.5 log units for diverse metal-ligand pairs. However, notable limitations persist, including sparse coverage of organometallic complexes and data at elevated temperatures (>100°C), which restricts applications in and .

Applications

Supramolecular and host-guest systems

In , stability constants quantify the strength of non-covalent interactions in host-guest systems, where a host encapsulates a guest through mechanisms such as hydrophobic effects, hydrogen bonding, and π-π stacking. These equilibria are crucial for designing molecular assemblies that mimic biological recognition without covalent bonds. The macrocyclic effect, briefly, contributes to enhanced binding in rigid hosts by reducing entropic penalties upon complexation. Host-guest equilibria are exemplified by inclusions, where β-cyclodextrin forms 1:1 complexes with various guests, exhibiting constants with log K typically in the range of 2-4, indicating moderate affinity driven by guest size complementarity and hydrophobic inclusion within the toroidal cavity. For instance, derivatives bind with log K₁ values around 3, as determined by UV-Vis , highlighting the role of van der Waals contacts in stabilizing the complex. In supramolecular polymers, amplifies overall , where the cumulative association constant β_n greatly exceeds the product of individual stepwise constants K_i due to templating effects that align subsequent monomers. This positive , often by factors exceeding 10^2, enables chain growth beyond statistical expectations, as seen in hydrogen-bonded arrays where initial binding nucleates extended structures. Representative examples include calixarene-metal complexes, where p-tert-butylcalixarene derivatives extract alkali metals with stability constants increasing from Cs⁺ to Na⁺ (log K up to ~4.5 for Na⁺), attributed to cavity size matching and ion-dipole interactions, facilitating selective separation in solvent extraction processes. Rotaxanes, mechanically interlocked molecules, demonstrate high stability with log K >10 for pseudorotaxane formation in tight fits, such as cyclic peptide threads with axles, where multiple non-covalent contacts yield K_a ≈ 10³–10⁵ M⁻¹, enabling persistent threading under physiological conditions. Preorganization in these systems enhances the overall stability constant β by up to 10³ through rigid host geometries that minimize conformational entropy loss, as quantified in multi-site receptors where cooperativity between binding sites boosts affinity beyond additive effects. These host-guest systems find applications in sensors, where changes in stability constants upon analyte binding modulate fluorescence or electrochemical signals; for example, calixarene-based hosts detect metal ions with selectivity dictated by log K differences exceeding 2 orders of magnitude. In the 2020s, self-assembling coordination cages have advanced catalytic applications, with Pd₂L₄ assemblies encapsulating guests at K ≈ 10⁶ M⁻¹ to promote reactions like hydrogenation by confining substrates in chiral environments, as revealed by NMR and ITC studies. Similarly, organic cages bind anions with attomolar affinity (K_a ≈ 10¹⁷ M⁻¹), stabilizing transition states for selective transformations in aqueous media.

Biological and medicinal contexts

Stability constants play a crucial role in understanding the binding of metal ions to biomolecules, particularly in metalloproteins where precise coordination ensures functional integrity. In heme-containing proteins like hemoglobin and myoglobin, the Fe^{2+} ion forms a stable complex with the porphyrin ring, enabling reversible oxygen binding while preventing unwanted oxidation. Similarly, in zinc-dependent enzymes such as carbonic anhydrase, Zn^{2+} is coordinated by three histidine imidazole nitrogen atoms and a water molecule (or bicarbonate), forming a tetrahedral complex with high stability that facilitates CO_2 hydration; the effective binding constant reflects the enzyme's reliance on hard-soft acid-base (HSAB) matching, where borderline Zn^{2+} pairs well with nitrogen donors. Copper in superoxide dismutase (Cu/Zn-SOD) exhibits even greater stability, with the binuclear Cu^{2+}-Zn^{2+} center essential for dismutation of superoxide radicals without releasing free Cu^{2+}. In medicinal applications, stability constants guide chelation therapy for heavy metal poisoning by prioritizing ligands that outcompete endogenous binders. For lead intoxication, ethylenediaminetetraacetic acid (EDTA) forms a highly stable Pb^{2+}-EDTA complex with \log \beta = 18.0, allowing urinary excretion of the toxic metal while minimizing redistribution to sensitive tissues like the brain. Deferoxamine, a siderophore-derived chelator, binds Fe^{3+} with exceptional stability (\log \beta \approx 30.6), effectively treating iron overload in conditions like thalassemia by sequestering excess ferric ions and promoting their renal elimination. Drug design leverages these constants to optimize metal-based therapeutics and imaging agents. Cisplatin's anticancer activity stems from its stable monoaqua form binding to the N7 position of guanine in DNA, forming intrastrand cross-links with an effective stability driven by chelation (association constants on the order of 10^4-10^5 M^{-1} for initial binding), which distorts the helix and triggers apoptosis. In magnetic resonance imaging (MRI), gadolinium(III)-diethylenetriaminepentaacetic acid (Gd-DTPA) serves as a contrast agent due to its thermodynamic stability (\log \beta = 22.8), ensuring minimal free Gd^{3+} release in vivo while allowing T1 relaxation enhancement for clear visualization of tissues. Recent advances in targeted therapies incorporate metal into inhibitors to enhance selectivity and efficacy. A 2020 study on (III) prodrugs conjugated to EGFR inhibitors like demonstrated that metal-ligand stability, tuned via chelate design, enables intracellular activation and reduces off-target effects. In blood , iron involves competition between , which binds ^{3+} with high affinity (\log \beta \approx 20 under physiological conditions, accounting for carbonate synergy), and low-molecular-weight ligands like citrate (\log \beta \approx 11-15 for Fe^{3+}-citrate species), where dominates to prevent oxidative damage from non-transferrin-bound iron.

Environmental and analytical uses

Stability constants play a crucial role in by enabling the modeling of metal in natural waters, where ions can form complexes with toxic metals like mercury(II), influencing their and . For instance, the tetrachloro-mercurate(II) complex, HgCl₄²⁻, has an overall stability constant of log β₄ = 15.1 (at zero and 25°C), which promotes the formation of neutral or anionic that enhance mercury's uptake by aquatic organisms compared to the free Hg²⁺ . This shift, driven by concentrations in estuarine and coastal waters, increases mercury's to and , as chloride-bound forms penetrate cell membranes more readily than hydrolyzed . In soils and sediments, stability constants are essential for understanding metal binding to natural , such as humic and fulvic acids, which regulate mobility and . Conditional stability constants for Cu²⁺ binding to fulvic acid typically range from log K ≈ 10–12 (at environmental and ionic strengths), reflecting bidentate coordination through carboxylic and groups that immobilize and mitigate its in contaminated soils. These values, derived from potentiometric and spectroscopic measurements, highlight how ligands buffer metal release during or acidification events. Specific examples illustrate the environmental implications of complex stability. Pollution from EDTA, a strong chelator with log β ≈ 18 for Pb-EDTA, enhances lead(II) solubility and transport in , potentially mobilizing sorbed Pb from aquifers and increasing leaching risks at contaminated sites. Recent studies since 2020 have also examined (PFAS) interactions with metal ions, revealing that PFAS can form weak complexes (log K < 5) with cations like Cu²⁺ and Pb²⁺ in aqueous media, altering metal to soils and influencing co-contaminant transport in polluted ecosystems. Computational models like the Humic Aqueous Model (WHAM) integrate stability constants from curated databases to predict metal and toxicity in ecosystems, accounting for interactions with humics, inorganics, and competing ligands under varying ionic strengths typical of natural waters. In , stability constants guide the use of masking agents to selectively complex interfering ions during titrations, improving accuracy in multi-metal samples. For example, ions mask ³⁺ by forming the highly stable hexacyanoferrate(III) complex, [Fe(CN)₆]³⁻, with log β₆ = 31, preventing its interference in EDTA titrations of softer metals like Zn²⁺ or Cd²⁺ at alkaline pH. This approach exploits the vast difference in stability to isolate target analytes without physical separation. Furthermore, stability constants inform the design of ion-selective sensors, where ligands with tuned log K values (e.g., 10–15 for ) ensure high selectivity and sensitivity in detecting trace pollutants in environmental samples.

References

  1. [1]
    formation constant (F02485) - IUPAC Gold Book
    For a mononuclear binary complex, if a central atom (central group) M (the 'metal') and a ligand L have been defined, then in the following expressions K n ...Missing: coordination | Show results with:coordination<|control11|><|separator|>
  2. [2]
    Complex-Ion Equilibria - Chemistry LibreTexts
    Dec 21, 2023 · This constant may be called a stability constant or association constant; the units depend on the specific reaction it is describing. Common ...Introduction · Formation Constant · Example \(\PageIndex{2... · Chelation Effect
  3. [3]
    Stability Constants of Metal Complexes in Solution - IntechOpen
    Stability constant of the formation of metal complexes is used to measure interaction strength of reagents. From this process, metal ion and ligand interaction ...
  4. [4]
    COORDINATION CHEMISTRY
    The Keq of the reaction is the stability or formation constant. Large values of stability constants indicate stable complexes. If the equilibrium constant is ...
  5. [5]
    Quantum Chemical Approach for Calculating Stability Constants of ...
    Sep 28, 2018 · Stability constants are central to the multiscale modeling of the thermodynamic speciation, cycling, and transport of mercury (Hg) and other ...
  6. [6]
    The Relative Stability of Internal Metal Complexes. I ... - NASA ADS
    The values of stability constants of internal metal complexes in water-dioxan mixtures are greater than those in water. It is shown that inso~ubiity of internal ...
  7. [7]
    [PDF] Bjerrum J. Metal ammine formation in aqueous solution
    Besides giving a general treatment of equilibrium systems in which a metal ion stepwise bindsup to sixligands, this publication deals with the complex for-.
  8. [8]
    [PDF] NIST Critically Selected Stability Constants of Metal Complexes ...
    The NIST database contains critically selected stability constants of metal complexes. It includes information on ligand searching by name, formula, and ...
  9. [9]
    What's in a Name?—A Short History of Coordination Chemistry from ...
    Coordination chemistry enters this arena around 1704 when Diesbach accidentally prepared a new blue pigment (Figure 2a) which became known as Prussian blue ( ...
  10. [10]
    Brief History of the Thermodynamics of Complex Equilibria in Solution
    Nov 4, 1994 · The experimental and computational problems were solved in the 1940s mainly through the investigations of Jannik Bjerrum and Ido Leden. The ...
  11. [11]
    Metal Ammine Formation in Aqueous Solution: Theory of the ...
    Title, Metal Ammine Formation in Aqueous Solution: Theory of the Reversible Step Reactions ; Author, Jannik Bjerrum ; Edition, reprint ; Publisher, P. Haase, 1957.
  12. [12]
    Der Chelateffekt - Schwarzenbach - 1952 - Helvetica Chimica Acta
    The chelate effect has been defined as the stability difference between a chelate complex and the corresponding complex with simple ligands. Starting from a ...Missing: history | Show results with:history
  13. [13]
    680. Methods for computing successive stability constants from ...
    Methods for computing successive stability constants from experimental formation curves. H. Irving and (Mrs.)H. S. Rossotti. Abstract.
  14. [14]
    IUPAC Stability Constants Database
    Sep 21, 2006 · The definitive database of metal-complex stability constants. Stability constants are the key to an understanding of equilibria in solution.<|control11|><|separator|>
  15. [15]
    Time line for development of stability constant programs
    ### Chronological Summary of Stability Constant Calculation Developments (Early 19th Century to 1970s)
  16. [16]
    Subcommittee on Solubility and Equilibrium Data (SSED) - IUPAC
    In the mid-1970s a group of chemists and chemical engineers came together within IUPAC to exhaustively compile and critically evaluate reports of experimental ...
  17. [17]
    [PDF] Trends in Stepwise Constants | Dalal Institute
    It has been observed that log K values for K1, K2, K3 and K4 are 4.3, 3.6, 3.0 and 2.3, respectively. This regular decrease in stepwise stability constants can ...
  18. [18]
    Hydrolysis of cations. Formation constants and standard free ...
    Hydrolysis of cations. Formation constants and standard free energies of formation of hydroxy complexes.
  19. [19]
    [PDF] Aluminium hydrolysis constants - NECTAR COST
    Equilibrium constants for hydrolysis and associated equilibria in critical compilations ... 3 Al3+ + 4 H2O ⇌ Al3(OH)4. 5+ + 4 H+. –13.94. −14.06 ± 0.22.
  20. [20]
    [PDF] Critical evaluation of stability constants of metal complexes of ...
    Abstract: Available experimental data on stability constants of proton (hydron) and metal complexes for seven complexones of particular biomedical and environ-.<|control11|><|separator|>
  21. [21]
  22. [22]
    [PDF] GUIDELINES FOR THE EXTRAPOLATION TO ZERO IONIC ...
    2.7 Ionic strength corrections with SIT: complex formation equilibria ... A is the Debye-Hückel constant, which is temperature dependent as listed in ...
  23. [23]
  24. [24]
    [PDF] The IUPAC Stability Constants Database (SC-Database)
    Temperature dependence may be calculated from the van't Hoff equation, provided the ∆H value is known. Ionic strength dependence may be calculated from the ...
  25. [25]
    the formation of [Cm(SO4)n]3−2n (n = 1, 2, 3) complexes at T = 25 ...
    Contrarily, the temperature dependency of the log K03 is very well described by the linear van't Hoff equation, assuming ΔrC0p,m = 0. Thus, the thermodynamic ...<|control11|><|separator|>
  26. [26]
    [PDF] Thermodynamic and Kinetic Aspects of Metal Complexes
    Thermodynamic Stability is measured in terms of the stability constant (eqbm. constant) of the complex formation process. Formation of complexes may take place ...
  27. [27]
    the Basics: Developing Advanced Metal–Organic Frameworks Using ...
    Fundamental chemistry concepts including HSAB theory and kinetic energy help explain enthalpies of MOF formation and the use of higher temperatures for MOF ...
  28. [28]
    Thermodynamics of complex formation: DL-isoserine with nickel(II ...
    This stands for a remarkable entropy contribution to the stability of the complexes, coming from redistribution of water molecules involved in the process.
  29. [29]
    NIST Standard Reference Database 46
    Equilibrium constants are given for 25 unless other temperatures were used and an enthalphy value was not available for conversion to 25 C. Equilibrium data ...
  30. [30]
    The Chelate Effect - Advances in Chemistry (ACS Publications)
    Jul 22, 2009 · The translational entropy contribution to stability, originally described by Schwarzenbach and Adamson, is broadened to include entropy ...
  31. [31]
    Protonation and Cu 2+ complexation constants of EDTA and ...
    In literature, cumulative stability constants are usually reported as a logarithmic function (logβ), and this is the usual parameter employed to report and ...
  32. [32]
    substitution in complex ions - stability constants - Chemguide
    The overall stability constant for this (as log K) is 18.7. Another copper-based chelate comes from the reaction with EDTA. This also has a high stability ...
  33. [33]
    [PDF] Metal-Ligand Equilibria in Solution - Dalal Institute
    The values of stepwise equilibrium constants for the formation of a particular metal-complex decrease successively in most of the cases i.e. K1 > K2 > K3 ...
  34. [34]
    Macrocyclic effect on the stability of copper(II) tetramine complexes
    Redox metal–ligand cooperativity enables robust and efficient water oxidation catalysis at neutral pH with macrocyclic copper complexes.Missing: tren | Show results with:tren
  35. [35]
    The Solution Chemistry of Cu 2+ –tren Complexes Revisited ...
    Apr 4, 2012 · The stability constants and the UV/Vis and EPR spectra of the three species have been determined. The position of the absorption maximum clearly ...
  36. [36]
    Experimental and Computational Studies of the Macrocyclic Effect of ...
    Jan 23, 2009 · The macrocyclic [CuII(9-aneN3)His]•2+ complex has substantially higher intramolecular PT barriers than those of the [CuII(dien)His]•2+ ...
  37. [37]
    The Effect of pH on the Stability of EDTA Chelates - Wellyou Tech
    Sep 15, 2025 · However, under acidic conditions (pH < 4–5), their stability decreases significantly. Mechanism: In acidic solutions, hydrogen ions (H⁺) ...
  38. [38]
    Crystal Field Stabilization Energy - Chemistry LibreTexts
    Jun 30, 2023 · A consequence of Crystal Field Theory is that the distribution of electrons in the d orbitals may lead to net stabilization (decrease in energy) ...
  39. [39]
    NIST46 | NIST - National Institute of Standards and Technology
    Jan 29, 2013 · NIST Critically Selected Stability Constants of Metal Complexes: Version 8.0. This database has been discontinued.Missing: 1970s | Show results with:1970s
  40. [40]
    Order of Stability of Metal Complexes - Nature
    Order of Stability of Metal Complexes. H. IRVING &; R. J. P. WILLIAMS. Nature volume 162, pages 746–747 ...
  41. [41]
    [PDF] GUIDELINES FOR THE DETERMINATION OF STABILITY - iupac
    and A can be taken as the constant of the Debye-HUckel equation. The Davies Equation has been shown to represent the activity coefficients of multiply.<|control11|><|separator|>
  42. [42]
    Rigorous Potentiometric Determination of Metal Complexes Stability ...
    In the present work an undergraduate laboratory practice is described corresponding to both experimental and computational methods used in potentiometric ...
  43. [43]
    Linear plots for complexometric potentiometric titrations of mixtures ...
    The application of Gran plots to potentiometric complexation titrations of mixtures of two metals with a selective electrode for one of the metal ions is ...
  44. [44]
    Determination of stability constants of complexes from ...
    A method for the calculation of the stability constants of complexes from spectrophotometric measurements is presented.
  45. [45]
    Spectrophotometric determination of stability constants by equimolar ...
    complexes. Therefore, a method of stability constant deter- mination should be judged first of all according to its abili- ty to be used for stable complexes.
  46. [46]
    9.7: Spectrophotometric Studies of Complex Ions
    Jun 28, 2020 · The method of continuous variations, also called Job's method, is used to determine the stoichiometry of a metal-ligand complex. In this method ...
  47. [47]
    Study of competitive reactions by solvent extraction: Determination ...
    The methods available for the determination of the stability constants of non-extractable complexes by solvent extraction are reviewed.
  48. [48]
    Determining Stability Constants Using the AKUFVE Technique
    Aug 22, 2019 · Table 1. Stability constants obtained from fitting of the solvent extraction data from, with some of the experimental conditions listed below ...
  49. [49]
    NMR relaxation studies in solutions of transition metal complexes. VI ...
    Determination and comparison of stability constants of uranium(VI) and vanadium(V) glycine complexes.
  50. [50]
    EPR spectroscopy: A versatile tool for exploring transition metal ...
    Sep 25, 2024 · This review introduces the EPR spectra of transition metal complexes, offering a comprehensive theoretical foundation along with illustrative examples
  51. [51]
    Thermodynamic Measurements: Potentiometry, UV/Vis Spectrometry ...
    Dec 4, 2023 · Calorimetric titration uses the heat of reaction to determine the stability constants. Since most complexation reactions involve the heat of ...
  52. [52]
    Calorimetric investigation of the complex formation between ...
    According to the model, data analysis allows determina- tion of the stoichiometry, stability constant and molar enthalpy of their complex formation. The ...
  53. [53]
    Predicting the Stability Constants of Metal-Ion Complexes from First ...
    Aug 9, 2025 · In principle, experimental stability constants can be derived by using first principle methods such as DFT, as it can be calculated as a Gibbs ...
  54. [54]
    Prediction of stability constants of metal–ligand complexes by ...
    Jan 29, 2024 · One can evaluate the relative binding strength of a ligand (Lig) to a metal ion (M) by determining the metal–ligand complex (MLig) stability ...Missing: hydrolysis | Show results with:hydrolysis
  55. [55]
    Complexation of 3p-C-NETA with radiometal ions - ScienceDirect.com
    Aug 15, 2024 · The purpose of this study is to evaluate the complexation of Ga 3+ , Tb 3+ , Bi 3+ , and Ac 3+ radiometal ions with 3p-C-NETA using density functional theory.
  56. [56]
    NIST SRD 46. Critically Selected Stability Constants of Metal ... - PDR
    This version of NIST Standard Reference Database 46 is a major enhancement to this widely used database which provides comprehensive coverage of interactions.
  57. [57]
    A model to estimate stability constants of amino acid chelates with ...
    According to IUPAC criteria [4] stability constants are grouped in four categories with respect to standard error in their log K values: recommended (S.E. < 0. ...
  58. [58]
  59. [59]
    IUPAC Stability Constants Database (SC-Database)
    The IUPAC Stability Constants Database is the definitive database of metal-complex stability constants, providing access to 24000 references for 9800 ligands.Missing: compilation history<|control11|><|separator|>
  60. [60]
    SC-Database and Mini-SCDatabase
    It contains all significant stability constants and associated thermodynamic data published from 1887 to the present day, including all stability constants ...
  61. [61]
    Database of Stability Constants of Metal Complexes - GitHub Pages
    Sep 25, 2023 · Stability Constant Explorer is a simple search program for NIST SRD 46 database, “Critically Selected Stability Constants of Metal Complexes”.
  62. [62]
    Hyperquad2013 - Stability Constants
    Hyperquad2013 is a Windows™ application that can that can be used to derive equilibrium constants from experimental data. The data are usually obtained by ...
  63. [63]
    HySS Hyperquad Simulation and Speciation - Stability Constants
    Hyperquad simulation and speciation (HySS): a utility program for the investigation of equilibria involving soluble and partially soluble species.
  64. [64]
    Visual MINTEQ – A freeware chemical equilibrium model for the ...
    Visual MINTEQ is a freeware chemical equilibrium model for the calculation of metal speciation, solubility equilibria, sorption etc. for natural waters.Other software · Welcome to Visual MINTEQ · Download · Web appMissing: stability constants environmental
  65. [65]
    Potentiometric and Blood Plasma Simulation Studies of Nickel(II ...
    In calculating the pmi curves, use has been made of the ECCLES model [3, 4] of blood plasma to which formation constants determined in potentiometric study have ...Missing: software | Show results with:software
  66. [66]
    Acid Dissociation Constant Calculator | pKa Prediction Software
    Accurately predict acid dissociation constants (pKa) directly from structure. Predictions are based on a database of >26000 compounds.
  67. [67]
    [PDF] Deliverable 2.8: First results on the thermodynamic databases and ...
    For each new discrete phase, we can calculate the stability constant from the chemical potential of the of the non-ideal solid solution phases with the same ...
  68. [68]
    Machine Learning for Metal-Ligand Stability Constants
    Jul 25, 2022 · The stability constants of metal(M)-ligand(L) complexes are industrially important because they affect the quality of the plating film and ...
  69. [69]
    Review Where to find equilibrium constants? - ScienceDirect.com
    Nov 20, 2019 · This work surveys eleven major compilations of equilibrium constants including those from authoritative groups such as IUPAC, NIST, and NEA.
  70. [70]
    Dynamics of guest binding to supramolecular systems
    This chapter presents an overview of the techniques available to measure the dynamics of supramolecular systems, and in particular host–guest complexes.Missing: enhancing | Show results with:enhancing
  71. [71]
    Development of QSAR model for predicting the inclusion constants ...
    In this study, the logarithm of 1:1 α-cyclodextrin inclusion constants (log Kα) for 195 organic chemicals was collected. With this parameter as the endpoint, a ...
  72. [72]
    Stability constants of the inclusion complexes of β-cyclodextrin with ...
    Aug 7, 2025 · It has been shown that the derivatives studied form host-guest type complexes. The calculated stability constants, reported as log K1, were ...
  73. [73]
    Supramolecular assembly dynamics - PMC - NIH
    Supramolecular assemblies demonstrate cooperativity. This cooperativity affects both the stability of the cluster and the mechanism of its formation and ...Self-Assembly · Figure 1 · Host-Guest Chemistry And...Missing: constants | Show results with:constants
  74. [74]
    Extraction and complexation of alkali and alkaline earth metal ...
    May 26, 2015 · The stability constants of these complexes are given in Table 1. As can be concluded from the described experimental results, calixarene ...
  75. [75]
    Rotaxanes of Cyclic Peptides | Journal of the American Chemical ...
    High stability constants (Ka ≈ 103−105 M-1) have been reported for 1:1 metal and 1:1 and 1:2 protonated amino acid ester complexes of these cyclopeptides ...
  76. [76]
    Influence of non-covalent preorganization on supramolecular ...
    Mar 30, 2015 · Cooperativity between multiple interaction sites renders the overall stability of a supramolecular system different from the sum of the ...
  77. [77]
    Supramolecular and molecular capsules, cages and containers
    Oct 1, 2024 · In this tutorial review, an overview is provided into the design principles, synthesis, characterisation, structural facets and properties of coordination ...
  78. [78]
    Interchangeable utilization of metals: New perspectives on the ...
    The predicted complex stability of divalent metal cations is described by the Irving–Williams series (Mg2+ < Mn2+ < Fe2+ < Ni2+ < Co2+ < Cu2+ > Zn2+) (34), ...
  79. [79]
    Impact of Zn, Cu, and Fe on the activity of carbonic anhydrase of ...
    The impact of zinc, copper, and iron on the duck erythrocyte carbonic anhydrase (CA) activity and the hemoglobin content in vitro culture were studied.Missing: stability metalloproteins SOD
  80. [80]
    Inhibition of enzymes by metal ion-chelating reagents. The action of ...
    Aug 10, 2025 · A value of K=8.0x10(-33)m(-2) has been found for the enzyme in equilibrium with 2 Cu(2+) ions (i.e. beta(2), the stability constant for diamine ...Missing: post | Show results with:post
  81. [81]
    Chelation in Metal Intoxication - MDPI
    The Pb-EDTA complex has high stability constant thus CaNa2EDTA was found to chelate lead from the body fluids, excreting PbNa2EDTA leaving Ca behind.
  82. [82]
    [PDF] Deferoxamine B: A Natural, Excellent and Versatile Metal Chelator
    May 28, 2021 · The high stability of Fe(III) complexes with DFOB is well known and is the basis of its biological activity as a siderophore and of its use in ...
  83. [83]
    Hybrid Gd 3+ /cisplatin cross-linked polymer nanoparticles enhance ...
    New hybrid nanoparticles permitted MRI monitoring of a cisplatin infusion while enhancing drug accumulation and DNA adduct formation in glioblastoma cells.
  84. [84]
    Stability constants for Gd3+ binding to model DTPA-conjugates and ...
    Five DTPA-amide and ester derivatives have been synthesized and their Gd3+ stability constants have been measured using a simple spectrophotometric method.Missing: design cisplatin DNA
  85. [85]
    Improving the Stability of EGFR Inhibitor Cobalt(III) Prodrugs
    Nov 21, 2020 · Critical evaluation of equilibrium constants involving acetylacetone and its metal chelates. ... Metal-Tyrosine Kinase Inhibitors: Targeted metal ...
  86. [86]
    Human transferrin: An inorganic biochemistry perspective
    Dec 15, 2021 · Under the serum like conditions of carbonate concentration (~27 mM) the binding constants are estimated to be log K'1 = 22.8 and log K'2 = 21.5 ...
  87. [87]
    The iron(III) coordinating properties of citrate and α ...
    May 21, 2024 · Blood contains citrate at a level between 100 and 150 µM and it has been demonstrated to be a major ligand of nontransferrin-bound iron ( ...
  88. [88]
    Biotoxicity of mercury as influenced by mercury(II) speciation
    These results clearly demonstrate that chloride enhances the tox- icity of mercury and that the critical chloride:Hg molar ratio increases as the mercury ...
  89. [89]
    Stability constants for the complexation of copper(II) ions with water ...
    Stability constants for the complexation of copper(II) ions with water and soil fulvic acids measured by an ion selective electrode
  90. [90]
    Simulation of the mobility of metal-EDTA complexes in groundwater
    Jan 4, 2002 · Reactive transport simulations were conducted to model chemical reactions between metal−EDTA (ethylenediaminetetraacetic acid) complexes ...
  91. [91]
    Towards Solving the PFAS Problem: The Potential Role of Metal ...
    Jul 17, 2022 · This Review focuses on what role metal-organic frameworks (MOFs) may play in addressing the PFAS problem.<|control11|><|separator|>
  92. [92]
    Windermere Humic Aqueous Model (WHAM)
    The Windermere Humic Aqueous Model (WHAM7) simulates the chemical reactions that occur when metals enter soil and water systems.
  93. [93]
    Stability Constant and Potentiometric Sensitivity of Heavy Metal ...
    This study uses QSPR models to predict stability constants and potentiometric sensitivity of ligands with heavy metals, and designs new coumarin-like ligands ...