Open Access Article
Raúl Eduardo
Linares-Jiménez
a,
Björn
Drobot
a,
Robin
Steudtner
a,
Satoru
Tsushima
ab,
Dominik
Goldbach
a,
Thorsten
Stumpf
a,
Susanne
Sachs
a and
Jerome
Kretzschmar
*a
aHelmholtz-Zentrum Dresden-Rossendorf, Institute of Resource Ecology, Bautzner Landstraße 400, 01328 Dresden, Germany. E-mail: j.kretzschmar@hzdr.de
bLaboratory for Zero-Carbon Energy, Institute of Science Tokyo, Tokyo 152-8550, Japan
First published on 12th November 2025
Although uranyl(VI)–malate systems have been repeatedly studied, there are still open questions regarding their structures, stoichiometry, and thermodynamic key parameters. We therefore examined the interactions between the uranyl(VI) ion, U(VI), and malic acid, H2Mal, using a multi-technique approach performing nuclear magnetic resonance spectroscopy (NMR), time-resolved laser-induced fluorescence spectroscopy (TRLFS), isothermal titration calorimetry (ITC), and ultraviolet-visible spectroscopy (UV-vis), complemented by density functional theory (DFT) calculations. In acidic solution (pH 1.5–5.5), by covering metal excess through ligand excess, two dinuclear complexes of 2
:
1 and 2
:
2 U(VI)
:
malate stoichiometry form predominantly. This species distribution is mainly influenced by the metal-to-ligand ratio given in solution. DFT and NMR confirmed that the 2
:
1 U(VI) malate complex involves a bridging hydroxo ligand (µ2-OH). In both the 2
:
1 and 2
:
2 complexes, malate features a (κ3O,O′,O″) coordination motif with two carboxylate groups and the alkoxylato group including bridging between two U(VI). In the 2
:
2 complex, changing the ligands’ relative orientation yields two geometric isomers. Thermodynamic quantities (ΔG, ΔH, and ΔS) and formation constants (log
β) of both complexes were determined by calorimetric titrations and TRLFS. The formation of the 2
:
1 and 2
:
2 species is endothermic and entropy-driven, with log
β of 17.1 ± 0.1 and 37.7 ± 0.1, respectively. Notably, even for U(VI) concentrations as low as 10 µM, dinuclear species are predominant, while mononuclear species exist only in very acidic and/or very dilute solutions. This study provides new data which complement and expand the understanding of both structures and thermodynamics of these complexes.
Malic acid has two stereoisomers, the L- and D-isomers, with only the L-isomer occurring naturally in biological systems.12 It has three functional groups, two carboxyl and one hydroxyl groups, with pKa values of 3.4 and 5.1 for the carboxyl groups13,14 and 14.5 for the hydroxyl group,15 and will thus hereafter be denoted as H2Mal. After consecutive deprotonation of the carboxyl groups, HMal− and Mal2− are formed. As for other hydroxycarboxylic acids, deprotonation of hydroxyl groups in aqueous solution occurs only under extreme conditions. However, as will be of importance in the complexation studies, H+ abstraction can be facilitated at much lower pH upon metal ion coordination (metal ion-promoted ligand deprotonation).16–18 Given the similarity in structural features and hence pKa values for the hydroxyl groups of citric acid and malic acid,15 deprotonation of the hydroxyl group of Mal2− during complexation with metal ions is anticipated and will be referred to as Mal−H3−.
Uranium (U) has a wide range of concentrations that can vary from 21.4 μg L−1 in surface waters to 749 μg L−1 in groundwater.19 However, these natural amounts of U are increased by anthropogenic activities, such as U mining and milling, military use, and phosphate industry – including fertilizers. In the nuclear industry, U is used as fuel for electricity production in nuclear power plants. It is considered a contaminant of concern due to its toxicity to different organisms such as plants, animals, and microorganisms.20–24 In humans, U has been observed to produce diseases in different tissues such as the brain, kidneys, liver, or lungs.25–27 U exposure to humans mainly occurs from food and water uptake.28,29 Although naturally occurring U has a low radiotoxicity, it shows a potent chemotoxicity.27 Thus, knowledge of the interaction and transfer behavior of U in the environment up to the food chain is key in radioecology and pivotal for risk assessments of radioactive contaminated sites as well as for developing effective remediation strategies to prevent serious health hazards for humans.
Malic acid has been identified as a notable chelating agent for a broad variety of metal ions.30–33 Therefore, interaction between U and H2Mal is expected when both are present in soil environments. Complexation of uranyl(VI) and UO22+ with H2Mal has already been the subject of several studies.34–39 Rajan and Martell reported the uranyl(VI) malate complex to be dinuclear, where each malate ligand is fully deprotonated (Mal−H3−) and shared between two U(VI) ions.36 Based on spectrophotometric studies in solution at pH 3.5, the formation of complexes having uranyl(VI)
:
malate ratios of 2
:
1 and 2
:
2 was suggested by Feldman and Havill.39 Allen et al. determined the structural parameters of a 2
:
2 uranyl(VI) malate dimer in solution between pH 2 and 4 by means of extended X-ray absorption fine structure spectroscopy (EXAFS).38 Only a limited number of studies have focused on the thermodynamic characterization of uranyl(VI) malate complexes. Feldman et al. reported on uranyl(VI) malate 2
:
2 dimers and 3
:
3 trimers.34 Rajan and Martell expanded the set of species by proposing 1
:
1, 2
:
2, 3
:
3, and 6
:
6 uranyl(VI) malate complexes by potentiometry with 1 M KNO3.36 Kirishima et al. inferred the formation of 1
:
1 and 1
:
2 uranyl(VI) malate complexes with potentiometry and isothermal microcalorimetry in 1 M NaClO4 at 25 °C.40 Although previous studies have provided thermodynamic data, several key questions remain unanswered. What spectroscopic or crystallographic evidence supports the assumed stoichiometries and proposed structures of the complexes? Given the fact that most studies date back several decades, combining technological improvements and insights from related systems, some species or structural features can now be verified or falsified. Correspondingly, what are the formation constants associated with the authentic species?
Nuclear magnetic resonance (NMR) is a powerful spectroscopic method for the determination of the structures of organic molecules and their complexes with metal ions. Hence, it has been used for clarifying the structures of some uranyl(VI) malate complexes. These studies revealed that these complexes are predominantly binuclear in nature, with the 2
:
2 uranyl(VI) malate complex being the predominant complex in an equimolar solution at low pH values (pH 2–4). In addition, a uranyl(VI) malate 2
:
1 complex was described, which is also present at low pH values.35,41,42 There are open questions regarding the structures (including stoichiometry and coordination modes) as well as stability constants of these complexes, which are the subject of the present work. Building upon former investigations, state-of-the-art NMR spectroscopic measurements can provide 1H and 13C NMR spectra with better spectral resolution at much lower concentrations. Modern techniques such as 17O NMR and calculations using density functional theory (DFT) provide deeper insights helping in elucidating possible structures. Determination of thermodynamic parameters including the formation constants (log
β) is carried out with two different approaches: time-resolved laser-induced fluorescence spectroscopy (TRLFS) and isothermal titration calorimetry (ITC). TRLFS enables the investigation of the species distribution, which can be used for the determination of the complex formation constants. ITC is performed to determine the formation constants and further thermodynamic parameters such as enthalpy (ΔH), entropy (ΔS), and Gibbs energy (ΔG) of the reaction. UV-vis spectroscopy is performed to verify the oxidation state of U and compare it with those of similar systems already studied, for instance uranyl(VI) citrate.
This work aims to gain a deeper understanding of uranyl(VI) complexation with malic acid by addressing both structural and thermodynamic aspects. Structural characterization focuses on dinuclear 2
:
1 and 2
:
2 complexes using NMR spectroscopy and DFT calculations. Thermodynamic data including formation constants and enthalpic/entropic contributions are reinvestigated through TRLFS and ITC. Examination of this specific system has the potential to offer novel insights into the general interactions between uranyl(VI) ions and hydroxycarboxylic acids.
NMR samples were prepared by the dilution of appropriate aliquots of stock solutions of UO2(NO3)2 and malic acid. Because of sample concentrations in the µM range, NMR solutions were prepared in pure D2O (99.95% D), and pD was adjusted with DCl (99% D) and NaOD (99% D), all purchased from Deutero. pD was corrected for deuterium according to the relationship pD = pH(read) + 0.4.43
For TRLFS experiments, uranyl(VI) malate samples were prepared with a U(VI) concentration of 50 µM in 0.1 M NaClO4 solution, where the ligand concentration was varied between 0 and 0.100 M, and adjusted to pH = 4.0. For cryo-TRLFS, the solutions were transferred into plastic cuvettes (Rotilabo disposable UV cuvettes XK26.1, Carl Roth), and then flash-frozen with liquid nitrogen and stored at −80 °C until measurement. TRLFS measurements were performed at −120 °C to avoid quenching effects.
ITC experiments were performed with 30, 50, and 100 µM U(VI) and 1.0 mM malic acid in 0.1 M NaClO4 at pH = 4.0.
UV-vis experiments were performed at pH = 3.0, with concentrations of uranyl nitrate ranging from 50 µM to 5.0 mM, and those of malic acid ranging from 0 through 50 mM, using a 1 mL quartz cuvette.
To prevent any potential light-induced sample degradation, the samples were covered with aluminum foil during their preparation.
![]() | (1) |
All TRLFS data were analysed using parallel factor analysis (PARAFAC) as the N-way toolbox in MATLAB R2020b software (The Mathworks Corporation),51 as described by Drobot et al.52
Fig. 1 depicts calculated structures of 2
:
1 uranyl(VI) malate complexes, all of which are characterized by the presence of a single malate ligand that serves as a bridge between two uranyl(VI) entities. The structure shown in (A) features a bridging hydroxo ligand (µ2-OH) and a (κ3O,O′,O″) motif including a bridging alkoxylato group of the malate ligand as proposed by Nunes et al.41 The other species (structures B through E) constitute two uranyl(VI) units being bridged only by malate in various ways. That is, the alkoxyl group coordinates exclusively with one uranyl(VI), while the C1 carboxyl group (structures B–D) or the C4 carboxyl group (structure E) binds concomitantly, resulting in the formation of five- or six-membered ring chelation motifs, respectively. The remaining carboxyl group coordinates the second uranyl(VI) either unidentately (structure D) or bidentately, thereby adjusting the number of water molecules coordinating to the second uranyl(VI) to three (plus one additional explicit second-shell water) or four. The latter configuration yields a hexagonal bipyramid (structures B and C, respectively). Relative to structure B, structures C, D, and E are less stable by +14.5, +7.7, and +18.5 kJ mol−1, respectively.
The structures of the 2
:
2 complex isomers have also been calculated. One isomer possesses a rotational axis perpendicular to the molecular plane while the other isomer possesses a rotational axis perpendicular to the O–U–O axes, corresponding to Fig. 2 structures A and B, respectively. Interestingly, the calculated Gibbs energies reveal that the difference between structures A and B of the 2
:
2 complex is only 1.9 kJ mol−1, slightly favoring the former. Taking into account the inherent uncertainties in the use of density functional theory calculations largely associated with the implicit solvation model,56 this result is in full agreement with the two isomers observable by the NMR experiment in a ratio of ∼ 7
:
3 (cf. Fig. S1–S5, SI). Just as a rough estimation of the magnitude of the energy contribution from carboxylate binding, the structure given in Fig. 2C reveals a penalty of about 49 kJ mol−1 upon dissociation of one uranyl(VI)–carboxyl oxygen bond.
NMR spectroscopy was performed in order to verify the species required to describe the thermodynamic model used in ITC. For the given concentration range and pH (pD) values, the aqueous speciation of U(VI) and malic acid can be described satisfactorily with three uranyl(VI) and three malate species, respectively: the free uranyl(VI) aquo ion and the free ligand as well as the two dinuclear complexes with 2
:
1 and 2
:
2 U(VI)
:
malate ratios, the latter of which is present as two geometric isomers35,41,42 that are invariant in their molar ratio (throughout 70
:
30). NMR spectra associated with the 2
:
1 and 2
:
2 complexes are depicted in Fig. 3 and 4.
![]() | ||
Fig. 3 Uranyl(VI) malate 2 : 1 complex: (A) 1H NMR spectrum showing the three 2 : 1 complex-associated signals of the CH group (C2) as well as that of the CH2 group (C3) possessing diastereotopic hydrogens with distinct resonances (a/b). The asterisk indicates the residual signal from solvent (HDO) signal suppression. (B) 13C{1H} NMR spectrum showing the four 2 : 1 complex-associated signals of the two carboxyl (C1 and C4) carbons as well as of the CH group (C2) and of the CH2 group (C3). (C) 17O NMR of UO2 oxygen atoms. The bottom spectrum is a reference spectrum showing signals of three aqueous uranyl(VI) species (U1–U3) with different degrees of hydrolysis; [U(VI)] = 58 mM, pD 3.5. U1, U2, and U3 refer to the UO22+ aquo ion, the dinuclear (UO2)2(OH)22+ and the trinuclear (UO2)3(O)(OH)3+ species, respectively.57 The middle spectrum was obtained from a D2O solution with [U(VI)] = 58 mM, [malate] = 17 mM, pD 3.5, with the corresponding four 2 : 1 complex-associated signals extracted from deconvolution shown as the top spectrum. Signal assignment in NMR spectra is according to the labeling stated with the structure. | ||
This set of spectra was recorded to unambiguously identify the species upon comparison with and to complement literature data by means of 17O NMR spectra. Since the 2
:
1 complex is of C1 symmetry, all sites are unique. Due to the complex's inherent asymmetry arising from the chirality of the ligand, the two uranyl(VI) entities exhibit distinct characteristics. That is, one uranyl(VI) unit is a part of a five-membered ring chelation motif while the other features a six-membered ring. Consequently, the two uranyl(VI) units, each bearing two “yl”-oxygens, give rise to four distinct signals (Fig. 3C).
In the case of the 2
:
2 complex, both isomers feature C2 symmetry, with the C2 rotational axes being perpendicular to the molecular plane or perpendicular to the O
U
O axes (Fig. 4), respectively. Correspondingly, in one isomer, each uranyl(VI) is part of a five- and a six-membered ring chelation motif. In the other isomer, one uranyl(VI) is part of two five-membered ring chelation motifs while the other uranyl(VI) is part of two six-membered ring chelation motifs, cf. Fig. 4A and B, respectively. Within one 2
:
2 complex isomer species, the two ligands involved are equivalent by symmetry. However, owing to the difference in overall symmetry the isomers behave as diastereomers, giving rise to individual sets of signals. Therefore, the NMR spectra show two sets of signals: one set of intense signals referring to the main isomer and one set of less intense signals referring to the minor isomer.
In principle, the ligand gives rise to three 1H and four 13C signals. Since the methylene protons are diastereotopic, they show distinct resonances, at δH around 2.7 and 2.8 ppm as well as between 3.5 and 3.7 ppm for the free and U(VI) bound ligand, respectively. Owing to the scalar spin–spin coupling between each other (geminal coupling, 2JH,H) and to the adjacent CH (vicinal coupling, 3JH,H), each signal appears as a more or less resolved doublet of doublets, showing signal splitting with a characteristic coupling constant (Karplus relationship). Upon complexation, the 1H signals shift downfield, especially that of the CH group is displaced by about 2.5 ppm (cf. Fig. S1, SI), which can be mainly attributed to the electron withdrawing effect of two uranyl(VI) entities (each with an effective charge of +3.2)58 being bridged (µ2-O) by the adjacent hydroxylate group.
Because of the reduced number of potentially overlapping lines, and the sensitivity of 1H NMR spectroscopy, the CH groups’ 1H signals are well-suited for the discrimination between species (Fig. S1, SI, black rectangle).
The 1H NMR spectra of the pD-titration series with concentrations analogous to those in ITC experiments, obtained at varying U(VI) to malate ratios, as well as those of the additional pD- and concentration-dependent series, are provided as Fig. S2 and S6, SI. All these spectra taken together reveal the ranges of existence of the species of interest from the ligand's perspective.
Although the CH signals of the 2
:
1 complex and the 2
:
2 complex's minor isomer are very close, viz. δH 6.99 and 6.98 ppm, respectively, taking into account that the most downfield signal (δH 7.03 ppm) refers to the major isomer, and both isomers’ signals appear invariantly in a 7
:
3 intensity ratio, the spectral signature of all three complex species allows for reliable qualitative and quantitative assignment. Considering that in the 2
:
1 and 2
:
2 uranyl(VI) malate complexes one and two ligand molecules contribute to the resulting signals, for excess U(VI), the relative fractions of the 2
:
1 vs. 2
:
2 species (both isomers) are 0.66 vs. 0.34 at pD 3.0 and 0.91 vs. 0.09 at pD 5.0, respectively (Fig. S2A, SI). In the case of equimolar (100 µM) U(VI) and malic acid, the relative fractions (2
:
1 vs. 2
:
2 species) change from 0.56 vs. 0.44 at pD 3.0 to 0.82 vs. 0.18 at pD 5.0 (Fig. S2B, SI). For 10 µM U(VI) and tenfold ligand excess, at pD 5 the corresponding relative fractions amount to 0.07 vs. 0.93 (Fig. S2C, SI). Signal intensities in the pD 2.0 spectrum are too low for quantification. Notably, within the detection limits of NMR, regardless of composition, U(VI) complexation starts at pD 3.0, and even for U(VI) concentrations as low as 10 µM, the observed complex species are dinuclear. In the range 3.0 < pH/pD < 5.6 and ligand excess, the 2
:
2 species are predominant. Supplementary results are obtained by UV-vis spectrophotometry and are consistent among the other methods. Corresponding spectra are provided in Fig. S10 and S11, SI. Briefly, both the wavelength at the absorption maximum λ(εmax) and the maximum molar absorptivity εmax of the uranyl(VI) chromophore increase upon addition of malic acid, however, with significant differences for the given U(VI) to malate ratios. That is, spectral characteristics change from 9.3 M−1 cm−1 at 414 nm (uranyl(VI) nitrate pH 3 solution) to 30.4 M−1 cm−1 at 427 nm for the 2
:
1 uranyl(VI) malate complexes and to 80.5 M−1 cm−1 at 434 nm for the 2
:
2 uranyl(VI) malate complexes, respectively.
:
1 and 2
:
2 uranyl(VI) malate complexes. After the verification of the stoichiometry and structures of these complexes, complex formation constants and thermodynamic parameters of these complexes were determined using TRLFS and ITC series.
For the calculation of the thermodynamic parameters, eqn (2) and (3) were considered as the equilibrium reactions for the 2
:
1 and 2
:
2 complexes, respectively. These formulas take into account the deprotonation of the carboxylic groups and also of the hydroxyl group (Mal−H3−) as an essential part of the dinuclear structures of the 2
:
1 and 2
:
2 uranyl(VI) malate complexes.
![]() | (2) |
| 2UO22+ + 2Mal−H3− ⇌ (UO2)2(Mal−H)22− | (3) |
All samples were analyzed under identical experimental conditions, thereby fulfilling the trilinearity condition required for PARAFAC. The fluorescence data were evaluated using the N-Way Toolbox in MATLAB R2020b. Speciation was constrained to the previously identified uranyl(VI) malate complexes, as described elsewhere.59 The resulting emission spectra and species distributions are presented in Fig. 5, while the spectral features of the individual species are summarized in Table 1.
| Species | Luminescence emission bands (nm) | ||||
|---|---|---|---|---|---|
| Uranyl(VI) cation | 485 | 507 | 530 | 556 | (584) |
2 : 1 uranyl(VI) malate |
486 | 508 | 532 | 558 | (586) |
2 : 2 uranyl(VI) malate |
(476) | 503 | 525 | 549 | (574) |
This targeted speciation enabled the extraction of formation constants for the uranyl(VI) malate complexes. Based on reactions (2) and (3), we determined log
β values of 16.7 ± 0.1 for the 2
:
1 complex and 37.6 ± 0.1 for the 2
:
2 complex. Additional information regarding the TRLFS data is provided as Fig. S8, SI.
Malic acid (1 mM) was titrated into aqueous U(VI) solutions at pH 4 in 19 individual injections, using three different U concentrations (30 µM, 50 µM, and 100 µM). These variations were chosen to simultaneously ensure reproducibility and moderate shifts of the concentration windows explored. Background heat effects were corrected by subtracting a blank titration (malate into buffer without U(VI)). The ITC data were analyzed using the same speciation model as applied for the TRLFS data evaluation.
Fig. 6 illustrates representative results for the 100 µM U(VI) titration, including the raw thermogram (A), the integrated heat curve and model fit (B), and the calculated species distribution (C). The corresponding results for the other U(VI) concentrations (30 and 50 µM) are provided as Fig. S9, SI.
The extracted formation constants were log
β = 17.1 ± 0.1 for the 2
:
1 uranyl(VI) malate complex and log
β = 37.7 ± 0.1 for the 2
:
2 uranyl(VI) malate complex. Consistent with previous studies on the complexation of f-block elements with aminocarboxylates,17 both complexation reactions were endothermic, indicating entropy-driven processes. The entropy contributions are likely attributed to stripping of coordinated water molecules from both the metal ion and the ligand during complex formation. Gibbs free energies (ΔG) and entropy changes (ΔS) were calculated using eqn (4) and (5):
ΔG = −RT ln (K) | (4) |
| ΔG = ΔH − TΔS | (5) |
A summary of all thermodynamic parameters obtained is provided in Table 2.
:
1 and 2
:
2 uranyl(VI) malate complexes at 25 °C in 0.1 M NaClO4 aqueous solution
β)
β values were extrapolated to standard conditions (zero ionic strength, I = 0 M) using the Davis model.55 The complex formation constants from ITC are in very good agreement with those obtained using TRLFS. This agreement highlights the reliability of these log
β values from different independent experiments with same conditions (molar ratios, pH, and ionic strength).
:
1 and 2
:
2 uranyl(VI) malate complexes at 25 °C. Uncertainties are represented by the standard deviation of at least three independent experiments
| Complex species | log β a (TRLFS) |
log β0 b (TRLFS) |
log![]() βa (ITC) |
log β 0b (ITC) |
|---|---|---|---|---|
| a I = 0.1 M NaClO4. b Extrapolated to zero ionic strength using Davies equation with the following parameters: A = 0.5093 kg1/2 mol−1/2 and b = 0.3 kg1/2 mol−1/2.55 | ||||
2 : 1 uranyl(VI) malate |
16.7 ± 0.1 | 17.1 ± 0.1 | 17.1 ± 0.1 | 17.5 ± 0.1 |
2 : 2 uranyl(VI) malate |
37.6 ± 0.1 | 37.4 ± 0.4 | 37.7 ± 0.1 | 37.5 ± 0.4 |
:
1 complex as well as 3.940 and 3.915 Å for the two 2
:
2 complex isomers, are in excellent agreement with the experimental value, evidencing the proximity of the two uranyl(VI) entities caused by the (κ3O,O′,O″) motif including a bridging alkoxylato group of the malate ligand for both complexes and ruling out structures B through E in Fig. 1.
The uranyl(VI) 2
:
2 complexes of both malate and citrate are isostructural and reveal very much similar spectroscopic features.16 That is, the εmax at λ(εmax) is 80.5 M−1 cm−1 at 434 nm in the case of malate and 83 M−1 cm−1 at 435 nm in the case of citrate.61 In both cases, d(U–U) is determined as (3.93 ± 0.02) Å,38 and the uranyl(VI) oxygens give rise to distinct 17O NMR signals mirroring the involvement of two ligands and two uranyl(VI) entities.16 With reference to the citrate system, taking into account the alkoxyl group's pKa value of 14.5,15 the complex stability of the corresponding citrate complex amounts to a log
β of 34.2 ± 0.1 (in 0.1 M NaClO4) and thus is similar to the 37.6 ± 0.1 determined for the uranyl(VI) malate 2
:
2 complex. Both these stability constants are very large in their own right, and yet that of the malate complex is still 3 logarithmic units larger. Besides slightly more basic carboxyl groups and less steric demands in malate, we hypothesize that the kinetic effects in the citrate complexes are responsible for the weakening. That is, citrate's additional CH2COO(H) residue, by means of NMR line widths and exchange spectroscopy shown to be involved in intramolecular site exchange reactions,16 slightly reduces the overall stability. In contrast, the malate complexes lack such a feature since all of the ligand functional groups are involved in coordination.
Interestingly, the 2
:
1 complex not only contains the malate ligand bridging between two uranyl(VI) entities but also a bridging µ-OH ligand. In both the 2
:
1 and 2
:
2 complexes, the bridging of of two uranyl units by malate's hydroxyl group is the consequence of metal ion-promoted H+ abstraction. The same phenomenon is observed for coordinating water ligands being deprotonated at a pH far below the corresponding pKa (i.e. hydrolysis). Since under the given conditions (50 µM U(VI), pH 4) the presence of other hydroxo species such as UO2(OH)+ or (UO2)2(OH)22+ is negligible,62 one can speculate that in the first instance the bridging malate complexation eases the concomitant approaching of the two uranyl(VI) units. The latter subsequently facilitates proton abstraction from coordinating water yielding the bridging hydroxo ligand resulting in the [(UO2)2(µ2-OH)(Mal−H)]0(aq) complex (for the given U excess). This pathway can be sketched as the (fast/complete) transition from structure D to structure A in Fig. 1.
Feldman and Havill studied U(VI) and malate complexation in acidic solutions with a 10 mM U(VI) concentration. They found two distinct complexes with U/malate stoichiometric ratios of 1 and 2, respectively. The former species was predominant under equimolar and ligand excess conditions up to pH 4.8.39 In our work at pH 3.5, no complex with a U/malate stoichiometric ratio <1 (such as a 1
:
2 uranyl(VI) malate complex) was detected. Hence, under the conditions investigated, three distinct U(VI) species are observed, viz. uranyl(VI) aquo ions as well as 2
:
1 and 2
:
2 uranyl(VI) malate complexes. In a subsequent study at pH 3.5–11, Feldman et al.63 addressed the binding motifs and complex structures, concluding that tridentate ligand binding already occurred in acidic solution. Assuming that each ligand coordinates only one uranyl(VI) at a time and that alkoxyl groups do not deprotonate in acidic solution, they struggled to bring in accordance the complexes’ structures and the number of titrated H+, requiring the inclusion of two bridging hydroxide ions (µ2-OH) in their proposed structure. In a later study, Feldman et al.34 stated that within the pH range of 2–4, the speciation is dominated by binuclear tridentate complexes of 1
:
1 stoichiometry, with the proposed structure later proven to be correct.
From potentiometric titrations in the pH range 2–8, Rajan and Martell36 concluded that, in accordance with earlier studies, at pH 2–4 the complexes are binuclear, bridged by ligand functional groups. They erroneously assumed that the binuclear complex forms in a straight and simple reaction through dimerization of two mononuclear 1
:
1 complexes. It was hypothesized that, in the mononuclear complex, malate would act as a tridentate ligand, while the alkoxyl group would undergo deprotonation. This has resulted in an inaccurate depiction of the equilibrium thermodynamics and the associated equations. This inaccuracy arises from the consideration of the pKa values of the carboxyl groups, while the pKa value of the ligand's OH group is overlooked. Therefore, the reported stability constants are of little use.
A further study reported on potentiometric and calorimetric titrations performed at 1–10 mM U(VI), presumably at low pH (not stated).40 Although stating that 5-membered rings are the most stable binding motif, without any evidence of the complex structure but just by analogy for the sake of comparing ligands of the same carbon chain length, malate is depicted to coordinate (solely) by its two carboxyl groups, thereby forming a 7-membered ring. The determined log
K values are assigned to uranyl(VI) malate complexes of 1
:
1 and 1
:
2 stoichiometry. However, from our own observations along with those discussed above with binuclear complexes that are predominant, we consider the assignment and the accompanied stability constants questionable.
In principle, we do not dispute the existence of mononuclear 1
:
1 and/or 1
:
2 uranyl(VI) malate complexes. Given the numerous studies at low pH (and rather high concentrations), for uranyl(VI) complexation not only by malate but also by its related compounds tartrate and citrate, repeatedly evidencing the superior formation of binuclear complexes, we believe that mononuclear complexes have a narrow range of existence. That is, they exist in relevant (detectable) quantities only at sufficiently low pH or/and in very dilute solutions, both conditions preventing bridging between two uranyl(VI) entities. As dilution beyond the micromolar range hampers spectroscopic detectability, we acquired the NMR spectra of D2O solutions 10 mM in U(VI) and 250 mM in malate at pD 1.5 through 3.0 as well as a pD 3.5 spectrum for 10 µM U(VI) and 100 µM malate (Fig. S6 and S7, SI). The mononuclear complexes constitute coordination by the CH(OH)COOH moiety, i.e., a 5-membered ring chelation motif by the more acidic carboxyl group and the protonated alkoxyl group, leaving the CH2COOH residue unbound. This is inferred from the downfield shift of the CH(OH) 1H signal to ∼5.3 ppm (compared to the 4.6 ppm in the free ligand at a comparable pD but the ∼7.1 ppm of the dinuclear complexes), while the unbound CH2COOH is merely shifted (cf. Fig. S7, SI). The spectra indicate that the mononuclear species are (a) predominant only at the lowest pD values, (b) occur as two complexes, attributed to 1
:
1 and 1
:
2 uranyl(VI) malates, and (c) show an upfield shift of the 1H signals (best seen for CH(OH)) ascribed to the (commencing) deprotonation of the less acidic CH2COOH carboxyl group, corroborating that this residue is unbound. Due to the very small existence range, and limited accessibility with analytic methods caused by the overall low concentration, we consider the determination of trustworthy thermodynamic quantities (above all, stability constants) of mononuclear uranyl(VI) malate species truly challenging. Nonetheless, our observations underscore the dominance and relevance of binuclear species, already at U(VI) concentrations as low as 10−5 M.
:
1 complex is the predominant uranyl(VI) malate species as long as U(VI) is in excess, otherwise, i.e., for equimolar or malate excess conditions, both the free uranyl(VI) aquo ion and the 2
:
1 complex are displaced by the 2
:
2 uranyl(VI) malate complex. Even in the micromolar concentration range, complex species detected by TRLFS and NMR were essentially dinuclear.
Comprehensive NMR spectroscopic measurements were performed to study the structures and stoichiometry of the complexes in solution under various conditions to support the TRLFS and ITC data interpretation. The results show that the 2
:
2 uranyl(VI) malate complex is predominant over a broad pH range from 3 to 6 even at concentrations as low as 10 µM. The 2
:
1 uranyl(VI) malate was also detected, but at lower pH values (2–4). DFT calculations assisted in the interpretation of NMR spectra and helped to unambiguously assign the individual species. Thermodynamic data were determined by TRLFS and ITC. New binding affinities and stability constants for both complexes (2
:
2 and 2
:
1 complexes) in 0.1 M NaClO4 solutions were determined.
The assignment of the two 2
:
2 complex isomers to the obtained individual sets of NMR signals is not a trivial task and requires further systematic investigation, which is beyond the scope of this study but might be a subject of future work.
The present results enhance our knowledge of U(VI) interaction with malic acid, and significantly improve both the structural elucidation and the thermodynamic database for modeling radionuclide behavior in the biogeosphere. Essential for radioecological aspects, the data contribute to a more reliable prediction of the bioavailability and mobility of radionuclides in the environment in general and in the rhizosphere in particular, which can ultimately contribute to more reliable safety and risk assessments for humans and environment as well as help in refining remediation strategies.
Further inquiries can be directed to the corresponding author.
| This journal is © The Royal Society of Chemistry 2025 |