Next Article in Journal
Development of Abraham Model Correlations for Solute Transfer into the tert-Butyl Acetate Mono-Solvent and Updated Equations for Both Ethyl Acetate and Butyl Acetate
Next Article in Special Issue
Solvation Structure and Ion–Solvent Hydrogen Bonding of Hydrated Fluoride, Chloride and Bromide—A Comparative QM/MM MD Simulation Study
Previous Article in Journal / Special Issue
Structures of Hydrated Metal Ions in Solid State and Aqueous Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solvent Exchange around Aqueous Zn(II) from Ab Initio Molecular Dynamics Simulations

by
Adrian Malinowski
and
Maciej Śmiechowski
*
Department of Physical Chemistry, Faculty of Chemistry, Gdańsk University of Technology, Narutowicza 11/12, 80-233 Gdańsk, Poland
*
Author to whom correspondence should be addressed.
Liquids 2022, 2(3), 243-257; https://doi.org/10.3390/liquids2030015
Submission received: 25 August 2022 / Revised: 12 September 2022 / Accepted: 14 September 2022 / Published: 19 September 2022
(This article belongs to the Special Issue Hydration of Ions in Aqueous Solution)

Abstract

:
Hydrated zinc(II) cations, due to their importance in biological systems, are the subject of ongoing research concerning their hydration shell structure and dynamics. Here, ab initio molecular dynamics (AIMD) simulations are used to study solvent exchange events around aqueous Zn2+, for which observation in detail is possible owing to the considerable length of the generated trajectory. While the hexacoordinated Zn(H2O)62+ is the dominant form of Zn(II) in an aqueous solution, there is a non-negligible contribution of the pentacoordinated Zn(H2O)52+ complex which presence is linked to the dissociative solvent exchange events around Zn2+. The pentacoordinated Zn(II) has a much tighter hydration sphere and is characterized by a trigonal bipyramidal structure, in contrast to the usual octahedral symmetry of the hexacoordinated complex. In total, two full exchange events are registered in the analyzed trajectory. AIMD simulations on an adequate length scale thus provide a direct way of studying such solvent exchange events around ions in molecular detail.

Graphical Abstract

1. Introduction

The coordination of zinc(II) cation is of paramount importance in biological systems, as zinc is the active center in hundreds of metalloenzymes, most often catalyzing hydrolysis reactions [1]. The importance of zinc goes beyond the role of cofactor in proteins, as the zinc cation is also actively utilized for intracellular signaling and subject to elaborate homeostasis mechanisms in the cell [2].
Due to the particular importance of water as the solvent in biological context, the structure and dynamics of hydrated ions is a topic of special significance in physical chemistry of solutions [3,4,5]. The influence of ions on water is often expressed in the terms of structural effects exerted by ions on the solvent, leading to the distinction between structure making and breaking ions [5].
The hydration of Zn(II) is also the subject of extensive research concerning its solvation shell structure and solvent exchange dynamics. To date, numerous experimental techniques have been applied to study Zn2+(aq), including vibrational spectroscopy [6,7,8,9,10,11,12], extended X-ray absorption fine structure (EXAFS) [13,14,15,16,17], X-ray diffraction (XRD) [18,19,20,21,22], and neutron diffraction (ND) [23,24]. On the other hand, computational chemistry methods have also been used to obtain molecular-level data on zinc hydration. The static structure of aqueous complexes of Zn(II) has been comprehensively studied with quantum mechanical (QM) methods, using both wavefunction theory and density functional theory (DFT) [9,10,25,26,27,28,29,30,31,32,33]. Both structural and dynamical nature of the hydration phenomenona may be in turn investigated with molecular dynamics (MD) simulations with an increasing level of complexity. Zn(II) hydration has been studied with force field-based MD simulations [17,34,35,36,37,38,39,40,41], as well as quantum mechanics/molecular mechanics (QM/MM) MD simulations [42,43,44,45,46] and ab initio molecular dynamics (AIMD) simulations [47,48,49,50].
AIMD based on DFT is a particularly suited technique for studying aqueous ionic solutions [51,52]. Beginning with the pioneering study of Be2+(aq) [53], many metal ions in an aqueous solution have been studied using AIMD simulations [54,55,56,57,58]. The current availability of fast and approximate DFT methods, such as density functional tight binding (DFTB) allows for the generation of extensive data sets on solvated ions in molecular solvents, including water [59]. The particular advantage of AIMD is the ability to study intermolecular polarization effects that ultimately influence many dynamic properties of the studied systems [54,55,60].
Computational methods are also indispensable for elucidating solvent exchange mechanisms around ions [61,62]. Experimental methods, rather than providing a mechanism of the exchange reaction, can only help in estimating the rate of this process [24]. The details of H2O exchange around Zn(II) have been studied using MD simulations and static QM calculations [28,29,30,37]. Although AIMD simulations, combining the quantum description of electrons with the classical propagation of nuclei according to Newton’s laws of motion, seem ideal for this purpose, the rather long timescale of water exchange in the first hydration shell of Zn(II) has thus far hampered direct observation of this process using AIMD.
In this work, in order to bridge this gap and obtain reliable data on the structure of Zn2+(aq) and the dynamics of the solvent exchange process in its first hydration shell, we use extensive DFT-based AIMD simulations with a computational setup that is proven to be especially suitable for investigating liquid water. We hypothesize that the inability of hitherto equilibrium AIMD studies to register any solvent exchange around Zn(II) was a limitation of the short timescale of the simulations, rather than the flows in the method itself. We also aim to register the interplay of well-defined solvation complexes of Zn2+(aq) that participate in the solvent exchange process.

2. Computational Methods

We studied aqueous Zn(II) in a system composed of a single Zn2+ cation and 100 H2O molecules, which was contained in a cubic simulation supercell with applied periodic boundary conditions. The volume of the system was chosen to represent the experimental density of H2O at 298 K (997 kg m–3 [63]) coupled with the absolute standard partial molar ionic volume of Zn2+(aq) ( 34.9  cm3 mol–1 [64,65]), resulting in cell volume 14 . 33 3  Å3. The starting configuration was prepared using GROMACS (v. 2018.5) tools [66] by solvating a centrally placed Zn2+ cation with a solvent configuration obtained from a well-equilibrated AMOEBA water simulation [67].
The system was initially subject to energy minimization and subsequently equilibrated using classical MD simulations in the canonical ( N V T ) ensemble at T = 298.15  K performed using Tinker-HP package (v. 1.2) [68]. The revised AMOEBA force field for liquid water [69] in combination with the AMOEBA parameterization of Zn(II) [38] were applied. The ultimate configuration from this N V T trajectory was then used to initialize the actual AIMD simulation.
AIMD simulations were performed with the cp2k computational suite (v. 6.0) [70,71,72], employing the DFT-based representation of the electronic structure as implemented in the Quickstep electronic structure module in cp2k [73]. The revPBE exchange-correlation functional [74,75] coded in the libXC library [76] was used, as it provides excellent reproduction of numerous static and dynamic properties of liquid water and aqueous ionic solutions [77,78,79,80,81]. A mixed Gaussian atomic orbitals with plane waves (GPW) representation of the electronic structure is applied in cp2k [82], and we used a molecularly optimized short-ranged double-zeta (DZVP-MOLOPT-SR-GTH) basis set for atomic orbitals [83], coupled with the auxiliary plane wave expansion of the electron density up to a 500 Ry cutoff. Though heavily contracted and featuring few diffuse primitive atomic orbitals, the chosen basis set performs favorably when compared to a more expensive triple-zeta polarized basis set for the description of liquid water [84]. Only valence electrons were treated explicitly, while core electrons were represented by norm-conserving GTH pseudopotentials [85] parameterized for the PBE functional. Additionally, the smoothing of the electron density and its derivative on the spatial integration grid was applied (via keywords XC_SMOOTH_RHO NN50 and XC_DERIV NN50_SMOOTH in cp2k), as it was previously found to significantly improve the stability of the local energetics of liquid water [86]. Dispersion effects were included via two-body DFT-D3 empirical dispersion correction with zero damping term [87] with the cutoff set to 15 Å. The studied system has a non-zero total charge ( + 2 ) and the charge compensation is provided by an implicit neutralizing jellium background [88], as implemented in cp2k.
The system was first equilibrated for ∼35 ps in an AIMD simulation in the canonical ( N V T ) ensemble at T = 298.15 K using the canonical velocity rescaling (CSVR) thermostat [89] with the time constant set to 2000 cm 1 (≈16.67 fs). The simulation time step was set to 0.5 fs. After this equilibration period, the thermostat was removed and a production AIMD simulation in the microcanonical ( N V E ) ensemble was continued for 100 ps, with data collection every 2 fs. The centers of maximally localized Wannier functions (MLWFs) [90] were also computed and stored every 2 fs. Analyses were performed using in-house code and VMD (v. 1.9.4) [91], which also served as a visualization tool, along with gnuplot (v. 5.2) [92].

3. Results

The static structure of the hydration shells of aqueous Zn(II) is summarized by radial distribution functions (RDFs) for the Zn⋯O and Zn⋯H pairs, as depicted in Figure 1. The studied system is large enough to fully contain two hydration shells of Zn2+, as clearly seen in the Zn⋯O RDF that features a prominent shallow minimum with g ( r ) 0 separating the first hydration shell from the second one. At first glance, the running integration number of g ZnO ( r ) is very close to 6 along this minimum, implying a predominantly hexacoordinated Zn(II) ion. However, its exact value of 5.8 suggests that during as much as 20% of the trajectory Zn2+ might in fact be pentacoordinated. Detailed parameters of the analyzed RDFs are gathered in Table 1.
The first hydration shell of Zn2+ extends up to 3.03 Å and contains a non-integer number of water oxygen atoms. In order to determine the instantaneous coordination number (CN) of Zn(II), one could simply count O atoms found closer than the r min , 1 value, which would be tantamount to accepting the Heaviside step function as a selector function for H2O molecules. However, this results in a discontinuous CN trajectory and adopting a smooth selector function instead can help visualize transitions between different coordination states of Zn2+(aq), as previously proposed [50]. In this work, the coordinated solvent molecules are selected using a logistic cutoff function:
C N ( t ) = i = 1 N H 2 O 1 1 + e r ZnO , i ( t ) r 0 s ,
controlled by the cutoff radius r 0 and the sharpness parameter s. r ZnO , i ( t ) is the instantaneous Zn⋯O distance for the water molecule i. Adopting r 0 = 3.1 Å and s = 0.08 Å ensures a smooth transition between a fully coordinated H2O molecule in the first hydration shell, where C N i ( t ) 1 , and a non-coordinated molecule in the second hydration shell, where C N i ( t ) 0 , cf. Figure 1.
The time evolution of thus obtained smooth coordination number of Zn2+ is shown in Figure 2. As clearly seen, even though a continuous C N ( t ) definition is accepted, its value provides a clear-cut distinction between the hexa- and pentacoordinated states of aqueous Zn(II). The zinc cation spends most of the time in a Zn(H2O)62+ complex (about 81.2% of the trajectory). The pentacoordinated complex may be unanimously found in ∼16.3% of the trajectory, while we ascribe the rest to transitional structures or short-time existence periods of one of the two dominating structures that are too short for inclusion in the analysis.
Armed with the knowledge about the existence of two distinct coordination states of Zn2+(aq), we revisit Figure 1 in order to compare the Zn⋯O RDFs in the regions of existence of the hexa- and pentacoordinated complexes. As seen in Figure 3, there is only a slight difference between the RDF based on the entire trajectory and the one based only on the fragments where Zn(H2O)62+ is found, which is unsurprising given the predominance of the hexacoordinated Zn(II). More importantly, the Zn⋯O RDF restricted to the trajectory periods where Zn(H2O)52+ exists reveals a major compression of not only the first but also the second hydration shell of the zinc cation. The first maximum shifts to 2.07 Å, while the second one to 4.07 Å. Thus, the two distinct complexes are characterized with different hydration patterns extending up to the entire second hydration shell.
Unlike static geometry optimizations, AIMD simulations can provide ensemble-averaged structures. In order to find the preferred geometry of hexa- and pentacoordinated Zn2+(aq), we isolated the first shell complexes from the relevant trajectory fragments, rotated them to a common reference frame and averaged the geometry. The results can be seen in Figure 4. Zn(H2O)62+ forms a regular octahedral complex with the average Zn⋯O distance equal 2.14 Å. On the contrary, Zn(H2O)52+ is characterized by a trigonal bipyramidal arrangement of water ligands. The Zn⋯O distance for the three equatorial molecules is 2.04 Å, while the axial O atoms are located on average 2.14 Å from the metal cation. The geometries of the two complexes in Cartesian coordinates are available in the Supplementary Materials, see Figure S1.
While the positions of O atoms tend to be fixed in the Zn(II) solvation cage, the hydrogen atoms of the first shell water molecules are relatively free to bend from the coplanar arrangement with the Zn–O vector. We define the tilt angle, β , as the angle formed between the Zn–O vector and the dipole moment vector of the coordinated H2O molecule. The two-dimensional distribution of water molecule configurations in the ( r ZnO , β ) plane is shown in Figure 5. The distribution of angles is evidently bimodal, with a global maximum at 0°, indicating a coplanar arrangement of the H2O molecular plane and the Zn–O vector, and a local maximum at ca. 30°, reflecting the presence of a substantial population of water molecules with tilted H atoms. We note here that the latter tend to be located slightly farther away from the central atom.
The presence of two complexes in aqueous solution of Zn(II) has far-reaching implications for the water molecules in its hydration shell. Ions are known to polarize water and the extent of this polarization is related to the changes in infrared spectrum of solution with respect to bulk water [54,55]. In Figure 6 we compare the probability distribution functions of the length of molecular dipoles of H2O molecules in the entire system vs the ones in the first hydration shell of Zn2+(aq). Unlike the dipole moment in bulk water, which has a symmetric Gaussian distribution [54,55,60], the curve for the studied system is right-skewed due to the substantial polarization of the neighboring H2O molecules by the zinc cation. The distribution can be fitted to a split Pearson VII peak shape with a maximum located at 2.78 D. On the contrary, the distributions of dipole moments of water molecules found in the two complexes are symmetric and can similarly be fit to a Pearson VII analytical shape. It is found that H2O molecules in Zn(H2O)52+ are polarized more significantly (on average 3.71 D) than the ones in Zn(H2O)62+ (mean 3.43 D).
We finally would like to estimate the time frame of a single exchange event based on the trajectories of the participating water molecules. As seen in Figure 2, we register two major exchange events connected to a change in CN of Zn(II), at around 0–12 ps and again at 78–86 ps. Additionally, a short-lived occurrence of Zn(H2O)52+ is found at ca. 64 ps. In order to shed light on the molecular details of these exchange events, we plotted the time evolution of the Zn⋯O distance of involved H2O molecules, as presented in Figure 7.
The first event begins as soon as the production trajectory starts. The water molecule W47 dissociates from the initially present Zn(H2O)62+ at 0.72–0.75 ps. Hereafter, the Wn notation indicates a H2O molecule with an arbitrary index n in the trajectory. The solvent exchange process is obviously dissociative, as no new solvent molecule coordinates to the cation for an extended period of time. In this event, the released H2O moves to the second hydration shell, as its separation from Zn2+ increases to ca. 4 Å (cf. Figure 1). This results in a transient Zn(H2O)52+ complex that persists up to ∼8.8 ps, at which moment an interesting concerted mechanism is observed. Namely, W47 is kicked off to the weakly organized third hydration shell of the cation, at 6 Å and beyond, while another second shell H2O, specifically W12, coordinates to zinc(II), restoring the Zn(H2O)62+ complex. This, however, proves short-lived, as W12 dissociates already at ∼10.1 ps and returns to the second hydration shell. Since a similar phenomenon is observed in further events, vide infra, such attempted exchange events that ultimately fail to stabilize the newly formed complex are most probably a common occurrence in Zn2+(aq). Somewhat surprisingly, the entire exchange event is finalized by a third water molecule (W22) that is captured by Zn2+ not from the second, but rather directly from the third hydration shell, initially >6 Å away from the cation. The hexacoordinated complex is restored at ∼11.9 ps and remains stable for an extended period of time. Interestingly, the coordination of W22 is seemingly again assisted by W47, since the time of arrival of the latter in the third hydration shell coincides with the start of movement of the former towards the inner hydration shells.
The second exchange event is only an attempted escape of W9 from Zn(H2O)62+ at ca. 64 ps. However, no other H2O molecule tries to enter the first hydration shell of Zn(II) this time and ultimately, after transiently wandering off to the verge of the second shell, W9 is recaptured by the complex and the stable octahedral arrangement of ligands persists.
The third exchange event (and the second successful one) resembles the first one to a great extent. It begins with the dissociation of W77 at ca. 78.3 ps. This H2O molecule is released initially to the second hydration shell, but quickly travels further to ultimately stay at ca. 5 Å separation from zinc(II) for an extended period, which is the location of the shallow minimum between the second and the poorly defined third hydration shell of the cation (cf. Figure 1). Interestingly, this perturbation of the hydration shells leads to two concerted movements of water molecules towards the central complex, initially that of W40, quickly followed by W49. Similarly to the first exchange event, the intervening water molecule (here, W40) ultimately does not manage to become properly coordinated to the cation. Instead, it remains at the very edge of the first hydration shell, only sligthly perturbing the newly formed Zn(H2O)52+ complex. After its release back into the second hydration shell at ca. 81 ps, the pentacoordinated complex remains stable for almost 4 ps. Then, an interesting cascade of concerted events can be noticed. W77 returns to the second hydration shell, apparently swapping its place with W40, but at the same time W49 manages to get captured by zinc(II), reinstating the hexacoordinated complex. Again, concerted movement of H2O molecules seems to be crucial for the mutually assisted exchanges between solvation shells leading to coordination number change of Zn2+(aq).

4. Discussion

The coordination of Zn2+(aq) has been hitherto the subject of numerous experimental and computational studies, as reviewed in the Introduction. There is an overwhelming consensus that Zn(II) forms a hexahydrated complex with the structure of a regular octahedron (point group Oh for the ZnO6 core) [9,10,11,14,15,16,17,18,19,20,21,25,27,32,34,35,36,38,40,41,42,43,44,45,46,47,48,50,93]. Reports claiming predominantly pentacoordinated Zn2+(aq) do surface [28,29,31,33,49,94], but appear to suffer from a size effect, i.e., the undercoordination of Zn(II) occurs most commonly in small aqueous clusters and the Zn(H2O)62+ complex is uniformly restored in the bulk limit [32,40]. Most notably for the present study, the coordination number of Zn(II) in an aqueous solution found from AIMD simulations seems to be dependent on the computational protocol, with dispersion corrections deemed particularly necessary to stabilize the Zn(H2O)62+ complex [50].
The Zn⋯O distances found experimentally and in computational investigations have been compiled by several authors [3,9,15,23,42,46,48,95]. EXAFS measurements find r ZnO = 2.05 –2.13 Å [13,14,15,16,17], while a survey of reported XRD investigations leads to a possible range of r ZnO = 2.08 –2.17 Å [3,18,19,20,21,22]. The available AIMD simulations report also a considerable spread of r ZnO = 2.02 –2.18 Å [47,48,49,50] while a slightly narrower range ( r ZnO = 2.11 –2.16 Å) is reported from QM/MM MD simulations [42,43,44,45,46]. In comparison, our global average of the position of the first peak in g ZnO ( r ) (2.11 Å, see Figure 1 and Table 1) is generally in the middle of the reported distances.
The particular advantage of our analysis is the rigorous separation of the trajectory fragments with hexa- vs. pentacoordinated Zn(II), thanks to the smooth instantaneous coordination number, see Figure 2. This allowed us to prove the tightening of the hydration shells of Zn2+(aq) with the dissociation of one H2O from the complex. Our results for the most probable Zn⋯O distance in Zn(H2O)62+ vs. Zn(H2O)52+ (2.13 Å and 2.07 Å, respectively, cf. Figure 3) are in good agreement with earlier AIMD simulations, a similar separation of trajectory fragments was proposed (2.11 Å and 2.05 Å, respectively [50]). Static geometry optimizations also lead to a general conclusion that the water molecules are located closer to Zn(II) in Zn(H2O)52+ than in Zn(H2O)62+ [26,27,29,31,32]. The hexacoordinated complex always maintains the octahedral arrangement of water oxygens (although the inclusion of hydrogens lowers the point group to Th [9]), which is also found in our time-averaged structure, cf. Figure 4. More controversial is the preferred geometry of Zn(H2O)52+. Based on geometry optimizations, both trigonal bipyramid (point group D3h) [26,31,32] and square pyramid (point group C2v) [27,29,94] have been confirmed as minima of the potential energy surface. It appears though that the presence of second shell H2O molecules stabilizes the former structure [29,40,94]. The presence of external hydration shells is thus crucial to the observation of the time-averaged trigonal bipyramidal structure that we report here, see Figure 4. The discrepancy between the Zn⋯O distance to equatorial and axial molecules (2.04 Å and 2.14 Å, respectively) is similar to that found previously in an isolated cluster (2.05 Å and 2.13 Å, respectively [26]).
Notably for AIMD simulations, our considerable system size with 100 H2O molecules allows us to record the RDF for the entire second hydration shell, as well as for the weakly formed third shell, see Figure 1. Due to the structure-making nature of Zn2+(aq), the second shell is also accessible experimentally. Previous EXAFS results indicate a second shell of 12.2 O atoms at an average distance of 4.40 Å [16] or 11.2 O atoms at 4.10 Å [14]. The distance to second shell from XRD measurements equals 4.02–4.26 Å [20,21,22]. Results from AIMD and QM/MM MD simulations generally support these experimental findings, with second hydration shell composed of 13–15 molecules at an average distance 4.3–4.5 Å [42,43,48,50]. In comparison, our global average of the position of the second peak in g ZnO ( r ) (4.11 Å, see Figure 1 and Table 1) is generally in the lower limit of the reported distances, while the number of O atoms (15.2) indicates a fairly crowded second hydration shell of Zn(II). Most importantly, the lowering of CN of the cation results in a major compression of this shell, with the RDF peak shifted to 4.07 Å (see Figure 3) and at least two water molecules released to the bulk solution, as the number of O atoms in the second shell decreases to 13.
Earlier studies on monovalent ions found that they are strongly polarized by water, but in turn do not polarize the solvent significantly (in the sense of only slight alteration of H2O dipole moment) [54,55,60]. In strike contrast to these findings, the water molecules in the first hydration shell of Zn2+(aq) are heavily polarized by the electric field of the cation, see Figure 6. For all H2O molecules, we find an average dipole moment of 2.78 D, which is slightly lower than the value usually found in AIMD simulations of liquid water (∼3 D) [54,60]. However, mean dipole moment of liquid H2O for revPBE/revPBE0 functionals is indeed slightly lower and in agreement with the present findings [96,97]. While the increase of dipole moment of H2O near the cation was previously found in MD simulations with the polarizable Amoeba force field, as well as AIMD simulations [38,48], the clear advantage of our analysis is the ability to separate the two forms of hydrated Zn(II). In the hexacoordinated complex, the dipole moment of the first shell water molecules increases by 0.65 D with respect to the bulk. Dissociation of this complex leads to a further increase of the average dipole moment and it is almost 1 D greater than for non-coordinated H2O. Thus, the water molecules in Zn(H2O)52+ are evidently more heavily polarized than in Zn(H2O)62+. The 0.6 D increase of the mean dipole moment of first hydration shell H2O has been noted previously in AIMD simulations [48]. Simulations with the Amoeba force field indicate that the water molecules closest to the cation have their dipole moments raised by more than 1 D [38].
Computational methods are excellent tools for studying solvent exchange mechanism around solvated ions [62]. The experimental estimate of water residence time on Zn(II) from quasi-elastic neutron scattering is a rather broad range of 0.1–5 ns [24]. Assuming the actual residence time is near the lower limit, it is in principle accessible to longer MD simulations. Previously, only force field description of the potential energy surface could provide the possibilities to reach the required time scale. However, the exact formulation of Zn(II)–H2O interactions is crucial for enabling solvent exchange events and they were frequently not observed even in prolonged MD simulations [16,34,35,36]. On the other hand, a water residence time of 2.2 ns was found in an MD simulation with the Amoeba force field [38] and a rate constant for water exchange equal 4.1 × 10 8 s–1 was also obtained [37]. Most AIMD or QM/MM MD simulations to date suffered from length limitations and no exchange events were observed [42,43,44,45,46,47,48]. In stark contrast to these earlier results, we managed to register two full solvent exchange events in the first hydration shell of Zn(II). While it is impossible to quantitatively estimate the water residence time from such a small sample, we note that the starts of successful exchange events are separated by ca. 78 ps, so that our results are not in disagreement with the experimental estimates provided the true residence time is close to the lower limit proposed previously [24]. We note here that the solvent exchange may be studied using accelerated sampling, e.g., metadynamics [37,50,93]; however, while such an approach provides the details of the potential energy surface of the hydrated cation, only equilibrium simulations can capture the actual timescale of the observed events.
Unfortunately, the question of how the two complexes differ on the free energy scale cannot be answered directly from our equilibrium simulation. However, we can roughly estimate this free energy difference from the fractions of the trajectory attributed to the penta- and hexacoordinated structures as Δ G = R T ln ( 0.163 / 0.812 ) 4 kJ mol–1. On the one hand, this is high enough that the pentacoordinated structure might prove too elusive to be registered experimentally, e.g., with EXAFS, noting also that the average coordination number of Zn(II) (5.8) is within the typical uncertainty of structure determination methods from the value of 6.0, indicative of perfect hexacoordination. On the other hand, this is low enough (only 1.6 R T ) to allow relatively free interconversion between the two structures. We note here that the free energy difference between [UO2(H2O)5]2+(aq) and [UO2(H2O)4]2+(aq) from ab initio metadynamics is ∼3 kJ mol–1, with similarly inconclusive experimental results as in the case of Zn(II) [98].
Our results indicate that the solvent exchange around Zn(II) follows a dissociative mechanism, cf. Figure 7. The release of a single water molecule from the first hydration shell leads to a transient Zn(H2O)52+ complex, the lifetime of which may be tentatively estimated as 8–10 ps, based on the limited sample of events we registered. A similar timescale of dissociative exchange events was noted before by Fatmi et al. [43]. A similar mechanism of solvent exchange around hexacoordinated Zn(II) was also inferred from static QM calculations [29,30,62]. Interestingly, the dissociation of a water molecule does not lead to increased CN in the second shell. On the contrary, it undergoes compression and major structural changes, as evident from Figure 3. Consequently, the registered exchange events require concerted movement of assisting solvent molecules, cf. Figure 7. This cooperative movement of H2O molecules enables the capture of a newly coordinated water molecule from as far as the third solvation shell of Zn(II).

5. Conclusions

In this paper, the structure of the hydrated zinc(II) cation has been studied using extensive AIMD simulations. In contrast to numerous earlier MD investigations at different levels of theory, which frequently found only hexacoordinated zinc with fixed first hydration shell waters, we clearly identify the presence of both hexa- and pentacoordinated complexes of Zn2+. While the octahedral Zn(H2O)62+ remains the dominant form of Zn(II) in an aqueous solution, Zn(H2O)52+ is present in about 16% of the trajectory. The average structure of the latter is a trigonal bipyramid and its presence results in major structural changes reaching up to the second hydration shell and beyond. Radial distribution functions indicate major compression of the second hydration shell of the pentacoordinated complex and molecular dipole distributions show profound polarization of first shell H2O molecules that is also heavily dependent on CN.
Owing to the considerable length of the generated trajectory, we were able to register two full solvent exchange events around aqueous Zn2+. They undoubtedly follow a dissociative mechanism and are driven by concerted solvent movement that may feature molecules up to the third hydration shell of Zn(II). While, due to the limited sampling, we cannot quantitatively estimate the water residence time on the zinc cation, the separation of successful events by ca. 80 ps seems to be consistent with the lower limit of experimental residence times (about 0.1 ns). The single exchange event, accompanied by the presence of Zn(H2O)52+, seems to take about 8–10 ps.
We conclude that equilibrium AIMD simulations on an adequate length scale provide a direct way of studying solvent exchange events around aqueous ions in molecular detail. The smooth selector function for coordinated solvent is a powerful tool that allows the identification of solvent exchange events and their preferred (associative or dissociative) mechanism. The analysis framework presented here may be tremendously useful for other solvated ions and more complicated systems containing Zn(II), such as concentrated zinc halide solutions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/liquids2030015/s1, Figure S1: Cartesian coordinates of the structures from Figure 4.

Author Contributions

Conceptualization, methodology, supervision, writing—original draft preparation, M.Ś.; investigation, formal analysis, visualization, writing—review and editing, A.M. and M.Ś. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Calculations were performed at the Academic Computer Center in Gdańsk (TASK).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AIMDab initio molecular dynamics
CNcoordination number
DFTdensity functional theory
DFTBdensity functional tight binding
EXAFSextended X-ray absorption fine structure
GTHGoedecker–Teter–Hutter (pseudopotentials)
MDmolecular dynamics
MLWFmaximally localized Wannier function
NDneutron diffraction
QMquantum mechanics
QM/MMquantum mechanics/molecular mechanics
revPBErevised Perdew–Burke–Ernzerhof (functional)
XRDX-ray diffraction

References

  1. Coleman, J.E. Zinc enzymes. Curr. Opin. Chem. Biol. 1998, 2, 222–234. [Google Scholar] [CrossRef]
  2. Maret, W. Zinc Biochemistry: From a Single Zinc Enzyme to a Key Element of Life. Adv. Nutr. 2013, 4, 82–91. [Google Scholar] [CrossRef]
  3. Ohtaki, H.; Radnai, T. Structure and Dynamics of Hydrated Ions. Chem. Rev. 1993, 93, 1157–1204. [Google Scholar] [CrossRef]
  4. Bakker, H.J. Structural Dynamics of Aqueous Salt Solutions. Chem. Rev. 2008, 108, 1456–1473. [Google Scholar] [CrossRef]
  5. Marcus, Y. Effect of Ions on the Structure of Water: Structure Making and Breaking. Chem. Rev. 2009, 109, 1346–1370. [Google Scholar] [CrossRef]
  6. Kristiansson, O.; Eriksson, A.; Lindgren, J. Hydration of Ions in Aqueous Solutions Studied by Infrared Spectroscopy. II. Application. Acta Chem. Scand. A 1984, 38, 613–618. [Google Scholar] [CrossRef]
  7. Stangret, J.; Libuś, Z. Influence of Zn(II), Mn(II), and Mg(II) Cations on the Vibrational Spectra of Water in Aqueous Perchlorate Solutions. Spectrosc. Lett. 1988, 21, 397–412. [Google Scholar] [CrossRef]
  8. Å. Bergström, P.; Lindgren, J.; Sandström, M.; Zhou, Y. Infrared Spectroscopic Study on the Hydration of Mercury(II), Cadmium(II), and Zinc(II) in Aqueous Solution and in the Hexahydrated Perchlorate Salts. Inorg. Chem. 1992, 31, 150–152. [Google Scholar] [CrossRef]
  9. Rudolph, W.W.; Pye, C.C. Zinc(II) hydration in aqueous solution. A Raman spectroscopic investigation and an ab-initio molecular orbital study. Phys. Chem. Chem. Phys. 1999, 1, 4583–4593. [Google Scholar] [CrossRef]
  10. Rudolph, W.W.; Pye, C.C. Zinc(II) Hydration in Aqueous Solution: A Raman Spectroscopic Investigation and An ab initio Molecular Orbital Study of Zinc(II) Water Clusters. J. Solut. Chem. 1999, 28, 1045–1070. [Google Scholar] [CrossRef]
  11. Mink, J.; Németh, C.; Hajba, L.; Sandström, M.; Goggin, P.L. Infrared and Raman spectroscopic and theoretical studies of hexaaqua metal ions in aqueous solution. J. Mol. Struct. 2003, 661–662, 141–151. [Google Scholar] [CrossRef]
  12. Wei, Z.F.; Zhang, Y.H.; Zhao, L.J.; Liu, J.H.; Li, X.H. Observation of the First Hydration Layer of Isolated Cations and Anions through the FTIR-ATR Difference Spectra. J. Phys. Chem. A 2005, 109, 1337–1342. [Google Scholar] [CrossRef]
  13. Miyanaga, T.; Watanabe, I.; Ikeda, S. Amplitude in EXAFS and Ligand Exchange Reaction of Hydrated 3d Transition Metal Complexes. Chem. Lett. 1988, 17, 1073–1076. [Google Scholar] [CrossRef]
  14. noz Páez, A.M.; Pappalardo, R.R.; Marcos, E.S. Determination of the Second Hydration Shell of Cr3+ and Zn2+ in Aqueous Solutions by Extended X-ray Absorption Fine Structure. J. Am. Chem. Soc. 1995, 117, 11710–11720. [Google Scholar] [CrossRef]
  15. Kuzmin, A.; Obst, S.; Purans, J. X-ray absorption spectroscopy and molecular dynamics studies of Zn2+ hydration in aqueous solutions. J. Phys. Condens. Matter 1997, 9, 10065–10078. [Google Scholar] [CrossRef]
  16. D’Angelo, P.; Barone, V.; Chillemi, G.; Sanna, N.; Meyer-Klaucke, W.; Pavel, N.V. Hydrogen and Higher Shell Contributions in Zn2+, Ni2+, and Co2+ Aqueous Solutions: An X-ray Absorption Fine Structure and Molecular Dynamics Study. J. Am. Chem. Soc. 2002, 124, 1959–1967. [Google Scholar] [CrossRef]
  17. Migliorati, V.; Mancini, G.; Tatoli, S.; Zitolo, A.; Filipponi, A.; De Panfilis, S.; Di Cicco, A.; D’Angelo, P. Hydration Properties of the Zn2+ Ion in Water at High Pressure. Inorg. Chem. 2013, 52, 1141–1150. [Google Scholar] [CrossRef]
  18. Bol, W.; Gerrits, G.J.A.; Van Panthaleon Van Eck, C.L. The Hydration of Divalent Cations in Aqueous Solution. An X-ray Investigation with Isomorphous Replacement. J. Appl. Cryst. 1970, 3, 486–492. [Google Scholar] [CrossRef]
  19. Ohtaki, H.; Yamaguchi, T.; Maeda, M. X-Ray Diffraction Studies of the Structures of Hydrated Divalent Transition-Metal Ions in Aqueous Solution. Bull. Chem. Soc. Jpn. 1976, 49, 701–708. [Google Scholar] [CrossRef]
  20. Licheri, G.; Paschina, G.; Piccaluga, G.; Pinna, G. X-ray Diffraction Study of Aqueous Solutions of ZnSO4. Z. Naturforsch. 1982, 37, 1205–1210. [Google Scholar] [CrossRef]
  21. Radnai, T.; Palinkas, G.; Caminiti, R. X-Ray Diffraction Study on Hydration and Ion-Pairing in Aqueous ZnSO4 Solution. Z. Naturforsch. 1982, 37, 1247–1252. [Google Scholar] [CrossRef]
  22. Dagnall, S.P.; Hague, D.N.; Towl, A.D.C. X-ray Diffraction Study of Aqueous Zinc(II) Nitrate. J. Chem. Soc. Faraday Trans. 2 Mol. Chem. Phys. 1982, 78, 2161–2167. [Google Scholar] [CrossRef]
  23. Powell, D.H.; Gullidge, P.M.N.; Neilson, G.W.; Bellissent-Funel, M.C. Zn2+ hydration and complexation in aqueous electrolyte solutions. Mol. Phys. 1990, 71, 1107–1116. [Google Scholar] [CrossRef]
  24. Salmon, P.S.; Bellissent-Funel, M.C.; Herdman, G.J. The dynamics of aqueous Zn2+ solutions: A study using incoherent quasi-elastic neutron scattering. J. Phys. Condens. Matter 1990, 2, 4297–4309. [Google Scholar] [CrossRef]
  25. Mhin, B.J.; Lee, S.; Cho, S.J.; Lee, K.; Kim, K.S. Zn(H2O)62+ is very stable among aqua-Zn(II) ions. Chem. Phys. Lett. 1992, 197, 77–80. [Google Scholar] [CrossRef]
  26. Bock, C.W.; Katz, A.K.; Glusker, J.P. Hydration of Zinc Ions: A Comparison with Magnesium and Beryllium Ions. J. Am. Chem. Soc. 1995, 117, 3754–3765. [Google Scholar] [CrossRef]
  27. Lee, S.; Kim, J.; Park, J.K.; Kim, K.S. Ab Initio Study of the Structures, Energetics, and Spectra of Aquazinc(II). J. Phys. Chem. 1996, 100, 14329–14338. [Google Scholar] [CrossRef]
  28. Hartmann, M.; Clark, T.; van Eldik, R. Theoretical Study of the Water Exchange Reaction on Divalent Zinc Ion using Density Functional Theory. J. Mol. Model. 1996, 2, 354–357. [Google Scholar] [CrossRef]
  29. Hartmann, M.; Clark, T.; van Eldik, R. Hydration and Water Exchange of Zinc(II) Ions. Application of Density Functional Theory. J. Am. Chem. Soc. 1997, 119, 7843–7850. [Google Scholar] [CrossRef]
  30. Rotzinger, F.P. Mechanism of Water Exchange for the Di- and Trivalent Metal Hexaaqua Ions of the First Transition Series. J. Am. Chem. Soc. 1997, 119, 5230–5238. [Google Scholar] [CrossRef]
  31. Pavlov, M.; Siegbahn, P.E.M.; Sandström, M. Hydration of Beryllium, Magnesium, Calcium, and Zinc Ions Using Density Functional Theory. J. Phys. Chem. A 1998, 102, 219–228. [Google Scholar] [CrossRef]
  32. De, S.; Ali, S.M.; Ali, A.; Gaikar, V.G. Micro-solvation of the Zn2+ ion—A case study. Phys. Chem. Chem. Phys. 2009, 11, 8285–8294. [Google Scholar] [CrossRef] [PubMed]
  33. Cooper, T.E.; O’Brien, J.T.; Williams, E.R.; Armentrout, P.B. Zn2+ Has a Primary Hydration Sphere of Five: IR Action Spectroscopy and Theoretical Studies of Hydrated Zn2+ Complexes in the Gas Phase. J. Phys. Chem. A 2010, 114, 12646–12655. [Google Scholar] [CrossRef]
  34. Obst, S.; Bradaczek, H. Molecular Dynamics Simulations of Zinc Ions in Water Using CHARMM. J. Mol. Model. 1997, 3, 224–232. [Google Scholar] [CrossRef]
  35. Chillemi, G.; D’Angelo, P.; Pavel, N.V.; Sanna, N.; Barone, V. Development and Validation of an Integrated Computational Approach for the Study of Ionic Species in Solution by Means of Effective Two-Body Potentials. The Case of Zn2+, Ni2+, and Co2+ in Aqueous Solutions. J. Am. Chem. Soc. 2002, 124, 1968–1976. [Google Scholar] [CrossRef]
  36. Arab, M.; Bougeard, D.; Smirnov, K.S. Molecular dynamics study of the structure and dynamics of Zn2+ ion in water. Chem. Phys. Lett. 2003, 379, 268–276. [Google Scholar] [CrossRef]
  37. Inada, Y.; Mohammed, A.M.; Loeffler, H.H.; Funahashi, S. Water-Exchange Mechanism for Zinc(II), Cadmium(II), and Mercury(II) Ions in Water as Studied by Umbrella-Sampling Molecular-Dynamics Simulations. Helv. Chim. Acta 2005, 88, 461–469. [Google Scholar] [CrossRef]
  38. Wu, J.C.; Piquemal, J.P.; Chaudret, R.; Reinhardt, P.; Ren, P. Polarizable Molecular Dynamics Simulation of Zn(II) in Water Using the AMOEBA Force Field. J. Chem. Theory Comput. 2010, 6, 2059–2070. [Google Scholar] [CrossRef]
  39. Yu, H.; Whitfield, T.W.; Harder, E.; Lamoureux, G.; Vorobyov, I.; Anisimov, V.M.; MacKerell, A.D.; Roux, B. Simulating Monovalent and Divalent Ions in Aqueous Solution Using a Drude Polarizable Force Field. J. Chem. Theory Comput. 2010, 6, 774–786. [Google Scholar] [CrossRef]
  40. Jana, C.; Ohanessian, G.; Clavaguéra, C. Theoretical insight into the coordination number of hydrated Zn2+ from gas phase to solution. Theor. Chem. Acc. 2016, 135, 141. [Google Scholar] [CrossRef]
  41. Xu, M.; Zhu, T.; Zhang, J.Z.H. Molecular Dynamics Simulation of Zinc Ion in Water with an ab Initio Based Neural Network Potential. J. Phys. Chem. A 2019, 123, 6587–6595. [Google Scholar] [CrossRef] [PubMed]
  42. Mohammed, A.M.; Loeffler, H.H.; Inada, Y.; Tanada, K.; Funahashi, S. Quantum mechanical/molecular mechanical molecular dynamic simulation of zinc(II) ion in water. J. Mol. Liquids 2005, 119, 55–62. [Google Scholar] [CrossRef]
  43. Fatmi, M.Q.; Hofer, T.S.; Randolf, B.R.; Rode, B.M. An extended ab initio QM/MM MD approach to structure and dynamics of Zn(II) in aqueous solution. J. Chem. Phys. 2005, 123, 054514. [Google Scholar] [CrossRef] [PubMed]
  44. Brancato, G.; Rega, N.; Barone, V. Microsolvation of the Zn(II) ion in aqueous solution: A hybrid QM/MM MD approach using non-periodic boundary conditions. Chem. Phys. Lett. 2008, 451, 53–57. [Google Scholar] [CrossRef]
  45. Rega, N.; Brancato, G.; Petrone, A.; Caruso, P.; Barone, V. Vibrational analysis of x-ray absorption fine structure thermal factors by ab initio molecular dynamics: The Zn(II) ion in aqueous solution as a case study. J. Chem. Phys. 2011, 134, 074504. [Google Scholar] [CrossRef]
  46. Riahi, S.; Roux, B.; Rowley, C.N. QM/MM Molecular Dynamics Simulations of the Hydration of Mg(II) and Zn(II) Ions. J. Chem. Phys. 2013, 99, 1–9. [Google Scholar] [CrossRef]
  47. Fujiwara, T.; Mochizuki, Y.; Komeiji, Y.; Okiyama, Y.; Mori, H.; Nakano, T.; Miyoshi, E. Fragment molecular orbital-based molecular dynamics (FMO-MD) simulations on hydrated Zn(II) ion. Chem. Phys. Lett. 2010, 490, 41–45. [Google Scholar] [CrossRef]
  48. Cauët, E.; Bogatko, S.; Weare, J.H.; Fulton, J.L.; Schenter, G.K.; Bylaska, E.J. Structure and dynamics of the hydration shells of the Zn2+ ion from ab initio molecular dynamics and combined ab initio and classical molecular dynamics simulations. J. Chem. Phys. 2010, 132, 194502. [Google Scholar] [CrossRef]
  49. Liu, X.; Lu, X.; Wang, R.; Meijer, E.J. Understanding hydration of Zn2+ in hydrothermal fluids with ab initio molecular dynamics. Phys. Chem. Chem. Phys. 2011, 13, 13305–13309. [Google Scholar] [CrossRef]
  50. Ducher, M.; Pietrucci, F.; Balan, E.; Ferlat, G.; Paulatto, L.; Blanchard, M. van der Waals Contribution to the Relative Stability of Aqueous Zn(2+) Coordination States. J. Chem. Theory Comput. 2017, 13, 3340–3347. [Google Scholar] [CrossRef] [Green Version]
  51. Marx, D.; Hutter, J. Ab Initio Molecular Dynamics; Cambridge University Press: Cambridge, UK, 2009. [Google Scholar]
  52. Hassanali, A.A.; Cuny, J.; Verdolino, V.; Parrinello, M. Aqueous solutions: State of the art in ab initio molecular dynamics. Phil. Trans. R. Soc. A 2014, 372, 20120482. [Google Scholar] [CrossRef] [PubMed]
  53. Marx, D.; Sprik, M.; Parrinello, M. Ab initio molecular dynamics of ion solvation. The case of Be2+ in water. Chem. Phys. Lett. 1997, 273, 360–366. [Google Scholar] [CrossRef]
  54. Śmiechowski, M.; Forbert, H.; Marx, D. Spatial decomposition and assignment of infrared spectra of simple ions in water from mid-infrared to THz frequencies: Li+(aq) and F-(aq). J. Chem. Phys. 2013, 139, 014506. [Google Scholar] [CrossRef] [PubMed]
  55. Śmiechowski, M.; Sun, J.; Forbert, H.; Marx, D. Solvation shell resolved THz spectra of simple aqua ions – distinct distance- and frequency-dependent contributions of solvation shells. Phys. Chem. Chem. Phys. 2015, 17, 8323–8329. [Google Scholar] [CrossRef]
  56. Chaudhari, M.I.; Soniat, M.; Rempe, S.B. Octa-Coordination and the Aqueous Ba2+ Ion. J. Phys. Chem. B 2015, 119, 8746–8753. [Google Scholar] [CrossRef]
  57. Schienbein, P.; Schwaab, G.; Forbert, H.; Havenith, M.; Marx, D. Correlations in the Solute–Solvent Dynamics Reach Beyond the First Hydration Shell of Ions. J. Phys. Chem. Lett. 2017, 8, 2373–2380. [Google Scholar] [CrossRef]
  58. Giacobello, F.; Mollica-Nardo, V.; Foti, C.; Ponterio, R.C.; Saija, F.; Trusso, S.; Sponer, J.; Cassone, G.; Giuffrè, O. Hydrolysis of Al3+ in Aqueous Solutions: Experiments and Ab Initio Simulations. Liquids 2022, 2, 26–38. [Google Scholar] [CrossRef]
  59. Gregory, K.P.; Elliott, G.R.; Wanless, E.J.; Webber, G.B.; Page, A.J. A quantum chemical molecular dynamics repository of solvated ions. Sci. Data 2022, 9, 430. [Google Scholar] [CrossRef]
  60. Śmiechowski, M. Unusual Influence of Fluorinated Anions on the Stretching Vibrations of Liquid Water. J. Phys. Chem. B 2018, 122, 3141–3152. [Google Scholar] [CrossRef]
  61. Helm, L.; Merbach, A.E. Water exchange on metal ions: Experiments and simulations. Coord. Chem. Rev. 1999, 187, 151–181. [Google Scholar] [CrossRef]
  62. Rotzinger, F.P. Treatment of Substitution and Rearrangement Mechanisms of Transition Metal Complexes with Quantum Chemical Methods. Chem. Rev. 2005, 105, 2003–2037. [Google Scholar] [CrossRef] [PubMed]
  63. Kell, G.S. Density, Thermal Expansivity, and Compressibility of Liquid Water from 0 °C to 150 °C: Correlations and Tables for Atmospheric Pressure and Saturation Reviewed and Expressed on 1968 Temperature Scale. J. Chem. Eng. Data 1975, 20, 97–105. [Google Scholar] [CrossRef]
  64. Pan, P.; Tremaine, P.R. Thermodynamics of aqueous zinc: Standard partial molar heat capacities and volumes of Zn2+(aq) from 10 to 55 °C. Geochim. Cosmochim. Acta 1994, 58, 4867–4874. [Google Scholar] [CrossRef]
  65. Marcus, Y. The Standard Partial Molar Volumes of Ions in Solution. Part 4. Ionic Volumes in Water at 0–100 °C. J. Phys. Chem. B 2009, 113, 10285–10291. [Google Scholar] [CrossRef] [PubMed]
  66. Abraham, M.J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J.C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1–22, 19–25. [Google Scholar] [CrossRef]
  67. Śmiechowski, M. Anion–water interactions of weakly hydrated anions: Molecular dynamics simulations of aqueous NaBF4 and NaPF6. Mol. Phys. 2016, 114, 1831–1846. [Google Scholar] [CrossRef]
  68. Lagardère, L.; Jolly, L.H.; Lipparini, F.; Aviat, F.; Stamm, B.; Jing, Z.F.; Harger, M.; Torabifard, H.; Cisneros, G.A.; Schnieders, M.J.; et al. Tinker-HP: A massively parallel molecular dynamics package for multiscale simulations of large complex systems with advanced point dipole polarizable force fields. Chem. Sci. 2018, 9, 956–972. [Google Scholar] [CrossRef]
  69. Laury, M.L.; Wang, L.P.; Pande, V.S.; Head-Gordon, T.; Ponder, J.W. Revised Parameters for the AMOEBA Polarizable Atomic Multipole Water Model. J. Phys. Chem. B 2015, 119, 9423–9437. [Google Scholar] [CrossRef]
  70. The cp2k Developers Group. cp2k v. 6.0, 2001–2018. Available online: http://www.cp2k.org/ (accessed on 13 September 2022).
  71. Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. cp2k: Atomistic simulations of condensed matter systems. WIREs Comput. Mol. Sci. 2014, 4, 15–25. [Google Scholar] [CrossRef]
  72. Kühne, T.D.; Iannuzzi, M.; Ben, M.D.; Rybkin, V.V.; Seewald, P.; Stein, F.; Laino, T.; Khaliullin, R.Z.; Schütt, O.; Schiffmann, F.; et al. CP2K: An electronic structure and molecular dynamics software package—Quickstep: Efficient and accurate electronic structure calculations. J. Chem. Phys. 2020, 152, 194103. [Google Scholar] [CrossRef]
  73. VandeVondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.; Hutter, J. Quickstep: Fast and accurate density functional calculations using a mixed Gaussian and plane waves approach. Comp. Phys. Commun. 2005, 167, 103–128. [Google Scholar] [CrossRef]
  74. Zhang, Y.; Yang, W. Comment on “Generalized Gradient Approximation Made Simple”. Phys. Rev. Lett. 1998, 80, 890. [Google Scholar] [CrossRef]
  75. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  76. Marques, M.A.L.; Oliveira, M.J.T.; Burnus, T. Libxc: A library of exchange and correlation functionals for density functional theory. Comp. Phys. Commun. 2012, 183, 2272–2281. [Google Scholar] [CrossRef]
  77. Marsalek, O.; Markland, T.E. Quantum Dynamics and Spectroscopy of Ab Initio Liquid Water: The Interplay of Nuclear and Electronic Quantum Effects. J. Phys. Chem. Lett. 2017, 8, 1545–1551. [Google Scholar] [CrossRef] [PubMed]
  78. Duignan, T.T.; Baer, M.D.; Schenter, G.K.; Mundy, C.J. Real single ion solvation free energies with quantum mechanical simulation. Chem. Sci. 2017, 8, 6131–6140. [Google Scholar] [CrossRef] [PubMed]
  79. Macchieraldo, R.; Esser, L.; Elfgen, R.; Voepel, P.; Zahn, S.; Smarsly, B.M.; Kirchner, B. Hydrophilic Ionic Liquid Mixtures of Weakly and Strongly Coordinating Anions with and without Water. ACS Omega 2018, 3, 8567–8582. [Google Scholar] [CrossRef] [PubMed]
  80. Ohto, T.; Dodia, M.; Imoto, S.; Nagata, Y. Structure and Dynamics of Water at the Water–Air Interface Using First-Principles Molecular Dynamics Simulations within Generalized Gradient Approximation. J. Chem. Theory Comput. 2019, 15, 595–602. [Google Scholar] [CrossRef] [PubMed]
  81. Ohto, T.; Dodia, M.; Xu, J.; Imoto, S.; Tang, F.; Zysk, F.; Kühne, T.D.; Shigeta, Y.; Bonn, M.; Wu, X.; et al. Accessing the Accuracy of Density Functional Theory through Structure and Dynamics of the Water–Air Interface. J. Chem. Theory Comput. 2019, 10, 4914–4919. [Google Scholar] [CrossRef]
  82. Lippert, G.; Hutter, J.; Parrinello, M. A hybrid Gaussian and plane wave density functional scheme. Mol. Phys. 1997, 92, 477–487. [Google Scholar] [CrossRef]
  83. VandeVondele, J.; Hutter, J. Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases. J. Chem. Phys. 2007, 127, 114105. [Google Scholar] [CrossRef] [PubMed]
  84. Galib, M.; Duignan, T.T.; Misteli, Y.; Baer, M.D.; Schenter, G.K.; Hutter, J.; Mundy, C.J. Mass density fluctuations in quantum and classical descriptions of liquid water. J. Chem. Phys. 2017, 146, 244501. [Google Scholar] [CrossRef] [PubMed]
  85. Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 1703–1710. [Google Scholar] [CrossRef] [PubMed]
  86. Jonchiere, R.; Seitsonen, A.P.; Ferlat, G.; Saitta, A.M.; Vuilleumier, R. Van der Waals effects in ab initio water at ambient and supercritical conditions. J. Chem. Phys. 2011, 135, 154503. [Google Scholar] [CrossRef]
  87. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  88. Durrant, T.R.; Murphy, S.T.; Watkins, M.B.; Shluger, A.L. Relation between image charge and potential alignment corrections for charged defects in periodic boundary conditions. J. Chem. Phys. 2018, 149, 024103. [Google Scholar] [CrossRef]
  89. Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101. [Google Scholar] [CrossRef]
  90. Marzari, N.; Vanderbilt, D. Maximally localized generalized Wannier functions for composite energy bands. Phys. Rev. B 1997, 56, 12847–12865. [Google Scholar] [CrossRef]
  91. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  92. Racine, J. gnuplot 4.0: A portable interactive plotting utility. J. Appl. Econom. 2006, 21, 133–141. [Google Scholar] [CrossRef]
  93. Brancato, G.; Barone, V. Free Energy Landscapes of Ion Coordination in Aqueous Solution. J. Phys. Chem. B 2011, 115, 12875–12878. [Google Scholar] [CrossRef] [PubMed]
  94. Cooper, T.E.; Carl, D.R.; Armentrout, P.B. Hydration Energies of Zinc(II): Threshold Collision-Induced Dissociation Experiments and Theoretical Studies. J. Phys. Chem. A 2009, 113, 13727–13741. [Google Scholar] [CrossRef] [PubMed]
  95. Marcus, Y. Ionic Radii in Aqueous Solutions. Chem. Rev. 1988, 88, 1475–1498. [Google Scholar] [CrossRef]
  96. Pestana, L.R.; Mardirossian, N.; Head-Gordon, M.; Head-Gordon, T. Ab initio molecular dynamics simulations of liquid water using high quality meta-GGA functionals. Chem. Sci. 2017, 8, 3554–3565. [Google Scholar] [CrossRef]
  97. Pestana, L.R.; Marsalek, O.; Markland, T.E.; Head-Gordon, T. The Quest for Accurate Liquid Water Properties from First Principles. J. Phys. Chem. Lett. 2018, 9, 5009–5016. [Google Scholar] [CrossRef]
  98. Atta-Fynn, R.; Bylaska, E.J.; de Jong, W.A. Free energies and mechanisms of water exchange around Uranyl from first principles molecular dynamics. Mater. Res. Soc. Symp. Proc. 2012, 1383, 113–118. [Google Scholar] [CrossRef]
Figure 1. Radial distribution functions, g ( r ) , for Zn⋯O (orange) and Zn⋯H (blue) pairs (solid lines, left ordinate), as well as their running integrals, N ( r ) (dashed lines, right ordinate). The black dotted line represents the smooth cutoff function used for water molecule coordination in Equation (1) (left ordinate).
Figure 1. Radial distribution functions, g ( r ) , for Zn⋯O (orange) and Zn⋯H (blue) pairs (solid lines, left ordinate), as well as their running integrals, N ( r ) (dashed lines, right ordinate). The black dotted line represents the smooth cutoff function used for water molecule coordination in Equation (1) (left ordinate).
Liquids 02 00015 g001
Figure 2. Time evolution of the smooth coordination number of Zn2+, C N ( t ) , along the AIMD production trajectory. Light green and light purple areas indicate regions assigned to hexa- and pentacoordinated aqueous Zn(II) complexes, respectively, while white regions indicate transitional periods excluded from detailed analysis.
Figure 2. Time evolution of the smooth coordination number of Zn2+, C N ( t ) , along the AIMD production trajectory. Light green and light purple areas indicate regions assigned to hexa- and pentacoordinated aqueous Zn(II) complexes, respectively, while white regions indicate transitional periods excluded from detailed analysis.
Liquids 02 00015 g002
Figure 3. Radial distribution functions, g ( r ) , for Zn⋯O pairs averaged over the entire AIMD trajectory (orange), as well as only over the regions of existence of hexacoordinated (green) and pentacoordinated (purple) aqueous Zn(II) complexes (solid lines, left ordinate), together with their running integrals, N ( r ) (dashed lines, right ordinate).
Figure 3. Radial distribution functions, g ( r ) , for Zn⋯O pairs averaged over the entire AIMD trajectory (orange), as well as only over the regions of existence of hexacoordinated (green) and pentacoordinated (purple) aqueous Zn(II) complexes (solid lines, left ordinate), together with their running integrals, N ( r ) (dashed lines, right ordinate).
Liquids 02 00015 g003
Figure 4. Time-averaged structure of (a) Zn(H2O)62+ and (b) Zn(H2O)52+ extracted from the respective regions of the AIMD trajectory.
Figure 4. Time-averaged structure of (a) Zn(H2O)62+ and (b) Zn(H2O)52+ extracted from the respective regions of the AIMD trajectory.
Liquids 02 00015 g004
Figure 5. The two-dimensional distribution function, g ( r ZnO , β ) , drawn on a logarithmic scale as log g ( r ZnO , β ) where β is the tilt angle of the coordinated H2O molecule as defined in the text.
Figure 5. The two-dimensional distribution function, g ( r ZnO , β ) , drawn on a logarithmic scale as log g ( r ZnO , β ) where β is the tilt angle of the coordinated H2O molecule as defined in the text.
Liquids 02 00015 g005
Figure 6. Normalized distribution functions of the molecular dipole moments of water molecules in the studied system, as well as water molecules in Zn(H2O)62+ (green) and Zn(H2O)52+ (purple).
Figure 6. Normalized distribution functions of the molecular dipole moments of water molecules in the studied system, as well as water molecules in Zn(H2O)62+ (green) and Zn(H2O)52+ (purple).
Liquids 02 00015 g006
Figure 7. Major successful and attempted exchange events registered during the AIMD production trajectory as depicted by the Zn⋯O distance of the participating water molecules. The Wn labels refer to the arbitrary index of the H2O molecule in the trajectory. Line colors indicate the leaving molecule (blue), the incoming molecule (red), as well as the intervening molecule (green). Black dotted line shows the location of the first minimum in the Zn⋯O radial distribution function (3.03 Å). Note the discontinuous abscissa axis.
Figure 7. Major successful and attempted exchange events registered during the AIMD production trajectory as depicted by the Zn⋯O distance of the participating water molecules. The Wn labels refer to the arbitrary index of the H2O molecule in the trajectory. Line colors indicate the leaving molecule (blue), the incoming molecule (red), as well as the intervening molecule (green). Black dotted line shows the location of the first minimum in the Zn⋯O radial distribution function (3.03 Å). Note the discontinuous abscissa axis.
Liquids 02 00015 g007
Table 1. Most important parameters of the radial distribution functions depicted in Figure 1: location of the first and the second maxima and minima and the respective running integration numbers.
Table 1. Most important parameters of the radial distribution functions depicted in Figure 1: location of the first and the second maxima and minima and the respective running integration numbers.
RDF r max , 1 / Å r min , 1 / Å N 1  1 r max , 2 / Å r min , 2 / Å N 2  2
Zn⋯O2.113.03 1 5.84.115.0715.2
Zn⋯H2.693.2112.04.755.7928.0
1N1 = N (rmin,1); 2 N2 = N (rmin,2) − N (rmin,2).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Malinowski, A.; Śmiechowski, M. Solvent Exchange around Aqueous Zn(II) from Ab Initio Molecular Dynamics Simulations. Liquids 2022, 2, 243-257. https://doi.org/10.3390/liquids2030015

AMA Style

Malinowski A, Śmiechowski M. Solvent Exchange around Aqueous Zn(II) from Ab Initio Molecular Dynamics Simulations. Liquids. 2022; 2(3):243-257. https://doi.org/10.3390/liquids2030015

Chicago/Turabian Style

Malinowski, Adrian, and Maciej Śmiechowski. 2022. "Solvent Exchange around Aqueous Zn(II) from Ab Initio Molecular Dynamics Simulations" Liquids 2, no. 3: 243-257. https://doi.org/10.3390/liquids2030015

Article Metrics

Back to TopTop