Vaska's complex is the air-stable, square-planar iridium(I) coordination compound with the chemical formula IrCl(CO)(PPh₃)₂, renowned for its ability to undergo reversible oxidative addition reactions with small molecules such as dioxygen, hydrogen, and halogens.[1] This 16-electron species features a central iridium atom bound to chloride, carbon monoxide, and two triphenylphosphine ligands in a trans arrangement, making it a prototypical example of coordinatively unsaturated organometallic chemistry.[2] First synthesized in 1961 by Lauri Vaska and John W. DiLuzio via the reaction of iridium trichloride with excess triphenylphosphine in a solvent providing carbon monoxide, such as dimethylformamide, the bright yellow crystalline solid is soluble in nonpolar solvents like benzene and chloroform.[1][3]The complex's reactivity stems from its low coordination number, enabling facile oxidative addition of substrates to the iridium center to form six-coordinate iridium(III) adducts; for instance, it binds dioxygen in a side-on (η²) peroxo fashion, shifting the CO stretching frequency from 1967 cm⁻¹ to higher values indicative of decreased back-donation from the metal.[2] This reversible oxygenation, first demonstrated in 1963, mimics the oxygen-carrying function of hemoglobin and marked one of the earliest synthetic models for biological dioxygen transport.[4] Similarly, it reacts with hydrogen to yield a dihydride and with HCl to form a chlorohydride, illustrating key mechanistic steps in catalytic processes.[1]Beyond fundamental studies, Vaska's complex has played a pivotal role in advancing homogeneous catalysis, serving as a precursor for iridium-based catalysts in hydrogenation of alkenes and as a benchmark for exploring ligand effects on reactivity.[3] Its discovery catalyzed broader interest in low-valent transition metal chemistry during the 1960s, influencing developments in Wilkinson's catalyst analogs and modern cross-coupling reactions.[3] The compound remains a staple in educational and research contexts due to its straightforward synthesis and versatile behavior.[2]
Structure and Properties
Molecular Geometry
Vaska's complex has the chemical formula \ce{IrCl(CO)(PPh3)2}, consisting of a central iridium atom coordinated to one chloride ligand, one carbonyl group, and two triphenylphosphine (PPh₃) ligands. The chloride and carbonyl ligands occupy trans positions in the coordination sphere, while the two PPh₃ ligands are arranged cis to each other.[5]The complex exhibits square planar coordination geometry, which is typical for 16-electron d⁸ transition metal complexes such as those of Ir(I). This arrangement is favored due to the low-spin electron configuration of the d⁸ metal center, resulting in a diamagnetic ground state with all electrons paired.[6]X-ray crystallographic analysis confirms the square planar structure, with key bond lengths including Ir–P at 2.330(1) Å, Ir–Cl at 2.382(3) Å, and Ir–C (of CO) at 1.791(13) Å. The bond angles are consistent with idealized square planar geometry, approaching 90° for adjacent ligands and 180° for trans pairs, though slight distortions arise from the steric bulk of the PPh₃ groups.[5]The electronic structure features a 16-electron count, achieved through sigma donation from the chloride, carbonyl, and phosphine ligands to the iridium center, supplemented by pi-backbonding from the filled d orbitals of Ir(I) primarily to the π* antibonding orbital of the CO ligand. The PPh₃ ligands act predominantly as sigma donors with minimal pi-acceptor capability, stabilizing the low-spin configuration.[7]
Physical Characteristics
Vaska's complex appears as a bright yellow crystalline solid.[7] This compound is air-stable in the solid state when stored under inert conditions but exhibits sensitivity to oxidation in solution, where it reacts with dioxygen to form a reversible adduct.[8][9] Its square planar geometry contributes to overall stability as a 16-electron d⁸ species.[7]The complex decomposes upon heating at approximately 215 °C without a distinct melting point.[10] It is diamagnetic due to its low-spin electronic configuration and non-hygroscopic, facilitating straightforward handling in laboratory settings.[7]Vaska's complex shows good solubility in common organic solvents, including benzene, toluene, chloroform, and tetrahydrofuran, but is insoluble in water and ethanol.[7] As a potential irritant and toxic material containing heavy metals, it requires careful handling under inert atmosphere to prevent oxidative side reactions.[10]
Synthesis
Standard Preparation
The complex was first synthesized in 1961 by refluxing hydrated iridium(III) chloride with excess triphenylphosphine (PPh₃) in ethanol for 2 hours under inert conditions.[1] Subsequent optimized standard preparations involve heating IrCl₃·3H₂O with excess PPh₃ in high-boiling solvents such as 2-methoxyethanol (190 °C, 2 hours, 86% yield), ethylene glycol (190 °C, 7 hours, 75% yield), or diethylene glycol (240 °C, 2 hours, 76% yield).[11]The most common laboratory procedure, a variation of these routes, involves refluxing IrCl₃·3H₂O (1 equiv) with excess PPh₃ (5 equiv) in dimethylformamide (DMF) or phenol at 100–150 °C for 1–2 hours under an inert atmosphere (N₂ or Ar), often with a CO atmosphere to facilitate carbonylation.[12] The reaction proceeds according to the simplified equation:\text{IrCl}_3 + 2 \text{PPh}_3 + \text{CO} \rightarrow \text{IrCl(CO)(PPh}_3\text{)}_2 + \text{byproducts}PPh₃ acts as both the coordinating ligand and a reducing agent to generate the Ir(I) center from the Ir(III) precursor.[11]Upon completion, the hot reaction mixture is filtered to remove insoluble byproducts, and the filtrate is diluted with methanol or petroleum ether to precipitate the yellow crystalline product, which is collected by filtration and washed with cold methanol followed by diethyl ether.[12] For further purification, the crude solid is dissolved in a minimal volume of hot benzene, the solution is filtered while hot, and the product is reprecipitated by cooling or dilution with petroleum ether.[12]This method typically provides 60–80% overall yield on a lab scale (millimolar quantities) after recrystallization, making it scalable for synthetic applications while maintaining high purity.[12]
Variations and Modifications
Alternative routes to Vaska's complex involve the reduction of iridium(III) precursors such as IrCl₃·3H₂O with excess triphenylphosphine (PPh₃) in the presence of a carbon monoxide source like dimethylformamide under reflux conditions.[8] Another approach utilizes ligand exchange reactions starting from Wilkinson's catalyst analogs, such as the iridium(I) complex IrCl(PPh₃)₃, by treatment with CO gas to displace one PPh₃ ligand and form the trans-carbonyl configuration.Modifications of Vaska's complex frequently entail substitution of the chloride ligand with other halides, such as bromide, yielding trans-IrBr(CO)(PPh₃)₂ through analogous reduction of IrBr₃ with PPh₃ and CO in high-boiling solvents like 2-methoxyethanol.[13]Phosphine ligand variations replace PPh₃ with bulkier or electronically tuned analogs, for example, tris(p-tolyl)phosphine [P(p-tolyl)₃] to produce trans-IrCl(CO)[P(p-tolyl)₃]₂, or water-soluble phosphines like tris(3-sulfonatophenyl)phosphine (TPPTS) to afford trans-IrCl(CO)(TPPTS)₂; these are typically prepared by direct reduction of IrCl₃ in the appropriate solvent or by ligand exchange from the parent complex.[14]Post-2000 developments have introduced greener synthetic protocols for derivative complexes, including the use of water-alcohol mixtures as solvents for preparing water-soluble analogs like trans-[IrCl(CO)(PTA)₂] (PTA = 1,3,5-triaza-7-phosphaadamantane), achieved by refluxing the parent complex with excess PTA in ethanol-water blends to enhance environmental compatibility and solubility in aqueous media.[14] These modifications often result in yields of 40–60% for sterically demanding phosphine-substituted variants, lower than the 75–86% achieved in the standard preparation due to increased steric hindrance during ligand coordination.[11]
Reactivity
Reversible Dioxygen Binding
Vaska's complex, trans-[IrCl(CO)(PPh₃)₂], undergoes reversible oxidative addition with dioxygen to form the corresponding dioxygen adduct. The reaction proceeds as follows:\text{trans-IrCl(CO)(PPh}_3)_2 + \text{O}_2 \rightleftharpoons \text{IrCl(η}^2\text{-O}_2\text{)(CO)(PPh}_3)_2This equilibrium is established under mild conditions, with the forward reaction occurring readily at ambient temperatures in the presence of O₂, while the reverse reaction is induced by evacuation or mild heating.[4] The trans arrangement of the chloride and carbonyl ligands in the starting square-planar complex is preserved in the product.[15]The mechanism involves a concerted oxidative addition, converting the d⁸ Ir(I) center to a d⁶ Ir(III) species, with dioxygen coordinating in a side-on (η²) fashion as a peroxo ligand. This binding mode weakens the O–O bond, as evidenced by structural data showing an elongated O–O distance of 1.47 Å in the adduct.[15][16] Thermodynamic studies indicate an exothermic process with ΔH ≈ -71 kJ/mol and a negative ΔS ≈ -155 J/mol·K, reflecting the loss of translational entropy upon O₂ association; the equilibrium favors the adduct at low temperatures but shifts toward dissociation above 0 °C under reduced pressure.[17]The dioxygen adduct adopts an octahedral geometry, featuring the η²-O₂ ligand perpendicular to the Ir–Cl and Ir–CO bonds, with the two triphenylphosphine ligands in trans positions. This red-brown compound is monomeric and stable under an atmosphere of O₂ at low temperatures (<0 °C), but it dissociates quantitatively upon warming to room temperature in vacuo, regenerating the yellow starting complex.[15]The rate of O₂ addition exhibits temperature dependence, with activation barriers low enough for rapid equilibration at ambient conditions, and solvent effects play a notable role: the reaction proceeds faster in aromatic solvents like benzene compared to aliphatic or polar media, likely due to better solvation of the nonpolar starting complex and transition state.[18]
Other Oxidative Additions
Vaska's complex, a 16-electron square-planar Ir(I) species, exhibits versatility in undergoing oxidative addition with various electrophilic substrates beyond dioxygen, forming stable 18-electron octahedral Ir(III) products through concerted addition to the metal center. This process increases the oxidation state from +1 to +3 and expands the coordination sphere from four to six, often with cis stereochemistry for the added ligands. The reactions typically proceed via a two-coordinate migratory insertion or direct bond breaking at the Ir atom, influenced by the electronic and steric properties of the substrate.[19][20]A representative example is the oxidative addition of dihydrogen, which yields the cis-dihydrido complex \ce{IrH2Cl(CO)(PPh3)2}. This reaction is reversible under certain conditions but kinetically slower than dioxygen binding, with rate constants indicating second-order kinetics dependent on both the complex and H2 concentrations.[19]The addition of hydrogen chloride produces the chlorohydrido adduct \ce{IrHCl2(CO)(PPh3)2}, where the H and Cl ligands occupy cis positions, consistent with the general stereochemistry of polar substrate additions to d8 complexes. This product has been characterized spectroscopically, showing shifts in the CO stretching frequency due to the increased electron density demands on the Ir(III) center.[21]Alkyl halides, such as methyl iodide, react via an SN2-like mechanism involving nucleophilic attack by the Ir(I) center on the carbon atom, followed by iodide transfer. The initial product is \ce{Ir(CH3)ICl(CO)(PPh3)2}, as shown in the equation:\ce{IrCl(CO)(PPh3)2 + CH3I -> Ir(CH3)(I)Cl(CO)(PPh3)2}Kinetic studies reveal that this addition is first-order in both reactants, with rates moderated by solvent polarity.[19]Sulfur dioxide undergoes reversible oxidative addition to form an S-coordinated \ce{Ir(SO2)Cl(CO)(PPh3)2} adduct, where SO2 acts as a two-electron donor in an η¹-S mode, marking an early demonstration of SO2 activation by transition metals. This reaction highlights the complex's affinity for Lewis acidic gases, with the adduct stable under ambient conditions.[22]These additions are generally favored for small, electrophilic substrates, but the steric bulk of the triphenylphosphine ligands hinders reactions with larger or more sterically demanding molecules, limiting the scope to unhindered electrophiles.
Spectroscopic Characterization
Infrared Spectroscopy
Infrared spectroscopy provides key insights into the bonding and electronic structure of Vaska's complex, IrCl(CO)(PPh₃)₂, with the CO stretching frequency serving as a primary probe for metal-ligand interactions. In chloroform solution, the complex exhibits a strong ν(CO) band at 1967 cm⁻¹, significantly lower than the 2143 cm⁻¹ for free CO, reflecting strong π-backbonding from the electron-rich d⁸ Ir(I) center into the CO π* orbital.[23] This frequency is characteristic of the square-planar geometry and trans arrangement of the phosphine ligands, which enhance the electron density at iridium.Oxidative addition reactions cause diagnostic shifts in ν(CO) due to changes in the metal's oxidation state and backbonding ability. For the dioxygen adduct, IrCl(CO)(O₂)(PPh₃)₂, the band appears at 2015 cm⁻¹ in solution, indicating weakened backbonding upon formal oxidation to Ir(III) and competition from the η²-O₂ ligand for the metal's d-electrons.[24] Similarly, addition of H₂ forms the cis-dihydride IrH₂Cl(CO)(PPh₃)₂ with ν(CO) at approximately 2033 cm⁻¹ (observed for the D₂ analog), a higher frequency consistent with reduced electron density at the metal in the Ir(III) state.[25] These shifts highlight IR's utility in monitoring oxidative additions in iridium chemistry.Additional IR features include the Ir–Cl stretching mode at ~310 cm⁻¹ in the far-infrared region, confirming the chloride's presence as a σ-donor ligand, and characteristic phenyl vibrations from the PPh₃ groups between 1400 and 1600 cm⁻¹, attributed to C=C stretches and ring deformations. The ν(CO) value for Vaska's complex and related Ir/Rh systems functions as an analog to the Tolman electronic parameter, quantifying phosphine donor ability (e.g., PPh₃'s moderate σ-donation yields low ν(CO) compared to more electron-withdrawing ligands).IR spectra are routinely recorded in solution (e.g., CHCl₃ or benzene) to capture dynamic behavior, where ν(CO) shows solvatochromism: frequencies increase slightly in more electrophilic solvents due to hydrogen bonding or polarization effects on the metal center.[24] Solid-state spectra, obtained via KBr pellets or ATR, often display broadened or split bands from lattice interactions but confirm the solution values. For adducts like the O₂ complex, temperature-dependent measurements reveal reversible dissociation above ~0 °C, with the 2015 cm⁻¹ band intensity decreasing as equilibrium favors the parent complex.[24]
NMR and Structural Analysis
Nuclear magnetic resonance (NMR) spectroscopy provides valuable insights into the electronic environment and ligand equivalence in Vaska's complex, trans-IrCl(CO)(PPh₃)₂. The ³¹P NMR spectrum exhibits a single peak at δ 24.5 ppm (in C₆D₆), indicative of the two equivalent triphenylphosphine ligands in the trans arrangement; coupling to ¹⁹⁵Ir (I = 3/2, 38% abundance) may appear as a doublet with J ≈ 300 Hz.[26] This chemical shift reflects the strong σ-donor and π-acceptor properties of the phosphine ligands coordinated to Ir(I).[26]The ¹H NMR spectrum of the parent complex shows aromatic protons from the PPh₃ ligands in the range of δ 7.0–8.0 ppm, typically as multiplets (e.g., δ 7.96 (m, 12H, o-Ph), δ 7.03 (m, 18H, m- and p-Ph) in C₆D₆), with no signals attributable to metal-hydride species.[26] These resonances confirm the absence of protons directly bound to iridium in the square planar Ir(I) structure.X-raydiffraction studies confirm the square planar geometry of Vaska's complex, with trans Cl and CO ligands. The Ir–P bond length is 2.330(1) Å, Ir–Cl is 2.382(3) Å, and Ir–C(O) is 1.791(13) Å, with the C–O bond at 1.161(18) Å and Ir–C–O angle of 175.1(12)°.[27] The P–Ir–P angle is approximately 99°, influenced by the steric bulk of the PPh₃ ligands.[27] Early structural analyses noted disorder between Cl and CO, resolved in later high-precision studies to affirm the trans configuration.[27]In solution, the dioxygen adduct exhibits fluxional behavior, with variable-temperature NMR revealing dynamic processes associated with O₂ binding and release. These studies demonstrate reversible coordination, with coalescence temperatures indicating rapid exchange between bound and free O₂ states.[17]Extended X-ray absorption fine structure (EXAFS) analyses complement crystallographic data by probing local bond metrics in non-crystalline environments, while density functional theory (DFT) modeling accurately reproduces experimental Ir–P (≈2.33 Å) and Ir–Cl (≈2.38 Å) distances, aiding understanding of electronic effects on geometry.[28]
History and Development
Discovery
In the early 1960s, amid burgeoning interest in organometallic chemistry and the synthesis of stable low-valent transition metal complexes, Lauri Vaska aimed to develop iridium-based compounds to explore their reactivity. Vaska and John W. DiLuzio reported the preparation of chlorocarbonylbis(triphenylphosphine)iridium(I) by refluxing iridium trichloride with excess triphenylphosphine in 2-methoxyethanol, in a seminal 1961 communication to the Journal of the American Chemical Society.[1] The reversible binding to dioxygen was unexpectedly observed and reported in 1963.[4][11]This breakthrough built on prior explorations of group 9 metal-phosphine systems. In 1957, L. M. Vallarino described the formation of rhodium complexes incorporating triphenylphosphine ligands through reactions of rhodium halides with the phosphine, providing early evidence for the stability of such coordination environments.[29] Independently, in 1959, Maria Angoletta investigated iridium carbonyl derivatives obtained from iridium halides and carbon monoxide, reporting species with compositions akin to Vaska's complex, though Vaska later acknowledged being unaware of her findings at the time.[11]The compound received its trivial name, Vaska's complex, in recognition of Lauri Vaska's contributions to its identification and study of its oxidative addition chemistry, with the eponym first appearing in the literature by 1966. Initial characterization relied on infrared spectroscopy, which revealed a characteristic CO stretching frequency around 1968 cm⁻¹ indicative of the square-planar geometry, supplemented by elemental analysis confirming the empirical formula.[1][11]
Key Milestones and Recent Advances
In the 1960s and 1970s, Vaska's complex served as a cornerstone for mechanistic investigations into oxidative addition reactions, establishing fundamental principles in organometallic chemistry. James P. Collman and colleagues conducted detailed studies on ligand effects influencing the rates and stereochemistry of oxidative additions to iridium(I) centers, highlighting how electronic and steric factors modulate reactivity with substrates like H₂ and alkyl halides.[19] Concurrently, Geoffrey Wilkinson's pioneering work on rhodium and iridium phosphine complexes, which contributed to his 1973 Nobel Prize in Chemistry for homogeneous catalysis, drew parallels with Vaska's system to elucidate migratory aptitude and reductive elimination pathways in catalytic cycles.[30]During the 1980s and 2000s, research expanded to rhodium analogs of Vaska's complex, such as trans-[RhCl(CO)(PPh₃)₂], which exhibited similar square-planar geometry but reduced reactivity toward oxidative additions due to the smaller ionic radius of rhodium. Computational modeling advanced understanding of O₂ binding, with ab initio molecular orbital calculations revealing substituent effects on electron affinities, ionization potentials, and the peroxo versus superoxo nature of the Ir-O₂ adduct.[31]From the 2010s to 2025, commemorative reviews marked the 50th anniversary in 2011, reflecting on the complex's enduring role in transition metal reactivity and catalysis.[11] Recent advances include steric and electronic tuning of Vaska-type iridium complexes through phosphine ligand modifications, enhancing selectivity in O₂ activation and biomimetic modeling of dioxygen carriers like hemoglobin.[32] In 2023, a coordination polymer incorporating Vaska's complex units demonstrated reversible O₂ binding and heterogeneous catalysis for reductive enamine formation from amides, bridging molecular and materials chemistry.[33] Furthermore, variants achieved high turnover numbers exceeding 10,000 in hydrosilylation of tertiary amides to silyl hemiaminals, advancing efficient C-N bond manipulations. In 2024, machine learning approaches explored the chemical space around Vaska's complex to optimize H₂-activation barriers, and iridium-catalyzed reductive deoxygenation of esters was developed using Vaska's complex variants.[7][34][35] These developments have refined insights into biomimetic O₂ activation, emphasizing low-valent iridium's role in selective oxygenation without over-reduction.
Applications and Significance
Model in Organometallic Chemistry
Vaska's complex, with the formula IrCl(CO)(PPh₃)₂, exemplifies a prototypical 16-electron square planar d⁸ organometallic complex, serving as a foundational model for understanding the geometry and electronicstructure of such species.[8] Its air-stable nature and well-defined trans arrangement of ligands make it an ideal representative for d⁸ iridium(I) chemistry, where the low-spin configuration stabilizes the unsaturated electron count.[11]In educational settings, Vaska's complex is prominently featured in organometallic chemistry curricula to demonstrate core concepts, including ligand field effects and the stability of square planar geometries. For instance, its synthesis is detailed in laboratory manuals as a hands-on experiment to explore d⁸ coordination and phosphine ligand roles.[36] Textbooks often highlight its use to teach how bulky triphenylphosphine ligands enforce the trans configuration and influence overall reactivity.[11]As a research model, Vaska's complex enables detailed investigations into metal-ligand interactions, such as π-backbonding from the iridium d-orbitals to the π* orbitals of CO and PPh₃, which strengthens the Ir-C and Ir-P bonds while shifting the CO stretching frequency.[7] Structural studies reveal the trans influence, where the strong σ-donor/π-acceptor CO ligand weakens the trans Ir-Cl bond more than the weaker σ-donor Cl affects the trans Ir-CO bond, affecting substitution patterns and bond lengths.[31] Comparisons with the rhodium analog RhCl(CO)(PPh₃)₂ underscore iridium's higher stability and, under specific conditions, higher catalytic activity in processes like hydrogenation, with conversion rates of 40% for iridium versus 24% for rhodium.[11]The complex's significance extends to theoretical frameworks in catalysis, as its reversible oxidative additions—such as with O₂—illustrate pathways to 18-electron octahedral products, challenging strict adherence to the 18-electron rule and informing exceptions for strong-field ligands in low-oxidation-state metals.[20] This has shaped understandings of electron counting in reactive intermediates.[37] However, the steric bulk of the PPh₃ ligands, characterized by a cone angle of 145°, hinders extensions to polynuclear or highly substituted systems, limiting its direct applicability in advanced architectures.
Catalytic and Synthetic Uses
Vaska's complex and its derivatives have found applications in reductive catalysis, particularly for the transformation of tertiary amides into enamines via hydrosilylation. In a seminal approach, the complex catalyzes the reductive activation of amides using silanes such as 1,1,3,3-tetramethyldisiloxane (TMDS), enabling the formation of enamine intermediates under mild conditions with low catalyst loadings (typically 1 mol% or less), achieving turnover numbers exceeding 100 for various substrates. This method demonstrates high chemoselectivity, tolerating functional groups like esters and nitro compounds, and has been applied to the synthesis of complex tertiary amine building blocks and natural products.Recent developments extend these capabilities to heterogeneous systems through coordination polymers derived from Vaska's complex. A 2023 study reported a low-valent metal-organic framework (LVMOF) assembled from tetratopic phosphine linkers and Ir(I) nodes, serving as a crystalline analogue of the homogeneous complex. This material catalyzes the reductive formation of enamines from amides with comparable efficiency to its soluble counterpart, while offering recyclability over multiple cycles without loss of activity. The polymer's structure maintains the reversible O₂ binding characteristic of Vaska's complex, enhancing its utility in oxygen-sensitive transformations.In hydrosilylation reactions, Rh and Ir variants inspired by Vaska's complex have shown promise for alkene silylation. A 2025 report highlighted a two-dimensional Rh(I) LVMOF, structurally analogous to Vaska's complex, that selectively hydrosilylates olefins such as 1-octene using polymethylhydrosiloxane (PMHS), operating heterogeneously with high selectivity and mild conditions (room temperature, no additives).[38]Iridium derivatives of Vaska's complex have similarly been employed in reductive deoxygenation of esters via hydrosilylation, achieving efficient silylation with turnover numbers up to 500 under ambient pressures.[39]Beyond organic synthesis, Vaska's complex inspires biomimetic applications in O₂ activation, modeling enzymatic processes like those in hemoglobin for catalytic oxygenations. Its reversible side-on binding of dioxygen facilitates studies in selective oxidation catalysis, though practical implementations remain limited to proof-of-concept systems.[7] In materials science, immobilization of Vaska-type complexes in nanoparticles and MOFs has emerged, with a 2025 review emphasizing their role in stable, low-valent frameworks for gas storage and sensing, leveraging enhanced thermal stability and recyclability compared to homogeneous forms.[40] These heterogeneous variants address challenges in catalyst recovery, enabling up to 10 recycles with minimal leaching, while retaining the mild reactivity of the parent complex.[40] However, transitioning from homogeneous to heterogeneous systems often requires balancing activity with structural integrity to avoid deactivation.