Emma N. Antonio‡
*ab,
Luis Kitsu Iglesias‡
cd,
Thomas P. Chaney
a,
Yinke Jiange,
Mrinal K. Beraf,
Julian Marsc,
Kee Sung Han
g,
Nathan T. Hahn
h,
Sanja Tepavcevic
i,
Edward Maginn
e,
Yong Zhang
e and
Michael F. Toney
*acj
aMaterials Science and Engineering, University of Colorado at Boulder, Boulder, CO 80303, USA. E-mail: Michael.Toney@colorado.edu; emma.antonio12@imperial.ac.uk
bDepartment of Chemical Engineering, Imperial College London, London, SW7 2AZ, UK
cDepartment of Chemical and Biological Engineering, University of Colorado Boulder, Boulder, CO 80303, USA
dElectrochemical Energy Systems Laboratory, Department of Mechanical and Process Engineering, ETH Zurich, 8092 Zurich, Switzerland
eDepartment of Chemical and Biomolecular Engineering, University of Notre Dame, Notre Dame, Indiana 46556, USA
fNSF's ChemMatCARS, Pritzker School of Molecular Engineering, University of Chicago, Chicago, Illinois 60637, USA
gJoint Center for Energy Storage Research, Pacific Northwest National Laboratory, Richland, WA 99354, USA
hMaterial, Physical and Chemical Sciences Center, Sandia National Laboratories, Albuquerque, NM 87185, USA
iMaterials Science Division, Argonne National Laboratory, Lemont, Illinois 60439, USA
jRenewable and Sustainable Energy Institute (RASEI), University of Colorado at Boulder, Boulder, CO 80303, USA
First published on 30th June 2025
The presence of ion nanoclusters or aggregates in electrolytes strongly impacts the electrolyte physical and chemical properties. Herein, we investigate the nature of ion nanoclusters in aqueous zinc sulfate electrolytes for zinc-ion batteries, specifically their size distribution, composition and lifetime, and subsequent influence on transport properties. Using a combination of X-ray scattering, molecular dynamics (MD) simulations and forward scattering calculations, we reveal the presence of free ions and a distribution of nanoclusters composed of solvent-separated ion pairs, whose size distributions are concentration-independent. Transient interactions between sulfate ions and long-lived hydrated zinc ions produce nanoclusters with varying sizes, compositions and geometries. Pair distribution functions calculated from MD simulations show strong agreement with experimental X-ray total scattering measurements. Nanoclusters were identified from MD simulations and used to forward simulate small-angle X-ray scattering (SAXS) of these nanoclusters. By fitting our measured SAXS using these forward simulations, we determine the distribution of nanocluster sizes in the electrolyte. Transport calculations from MD simulations and experimental measurements show that while nanoclusters hinder ion transport, their short-lived and dynamic nature enables efficient ion diffusion. Determining the nature of these nanoclusters is essential for understanding their impact on transport, solvation, and interfacial chemistry, and guiding the rational design of electrolytes for energy storage, chemical separations and environmental science.
Broader contextElectrolytes are essential to batteries, fuel cells, electrochemical separations, and environmental remediation technologies, where their solvation structure and nanoscale organization dictate key properties such as ion mobility and interfacial reactivity. While traditional models describe electrolytes as collections of free ions or simple ion pairs, recent studies suggest the presence of nanoclusters—ion aggregates that play a significant role in electrolyte behavior. Understanding these structures is critical for designing next-generation electrolytes with enhanced performance across energy and environmental remediation. In this work, we investigate nanocluster formation in aqueous ZnSO4 electrolytes, which are relevant for aqueous zinc batteries, a sustainable, low-cost, and safe alternative to lithium-ion batteries. Using a new multi-modal approach that integrates X-ray scattering, molecular dynamics simulations, forward scattering calculations, and transport measurements, we provide a quantitative understanding of nanocluster size, composition, and structure and their influence on ion transport. Our findings reveal a highly polydisperse system of nanoclusters with diverse sizes, compositions, and structures, introducing short-range correlations that reduce ionic conductivity. Yet, their transient and dynamic nature enables continuous reorganization, enabling facile transport. These insights introduce a new framework for electrolyte design, offering a pathway to improved energy storage systems and electrochemical technologies critical for clean energy and sustainability. |
Metal sulfate electrolytes are crucial in a range of fields, including energy storage, electroplating, and water treatment. Here we focus on the local solvation in aqueous ZnSO4 electrolytes, which are a low cost and safe option for zinc-based batteries with zinc metal anodes.15,16 Zinc-ion batteries have high volumetric capacity (5851 mAh cm−3), low redox potential (−0.76 V vs. SHE) and use widely available materials, making them attractive for grid-scale energy storage.17–19 Understanding nanoclusters in battery electrolytes is important due to their proposed effect on electronic properties, and thus redox properties; transport (deviations from transport predictions due to strong ionic correlations); and/or surface interactions including the electrical double layer, desolvation processes and interfacial reactions.1,14 These phenomena are particularly relevant for multivalent-ion-based electrolytes, including other battery chemistries such as calcium and magnesium.20–22
Gaining a comprehensive understanding of ion pairs and nanoclusters is challenging due to limitations in experimental and computational techniques. Experimental methods that probe ion pairs and nm-scale structures and chemistries include total scattering (or X-ray pair distribution function (XPDF)), Fourier-transform infrared spectroscopy (FTIR), Raman spectroscopy, X-ray absorption spectroscopy (XAS), nuclear magnetic resonance (NMR), and dielectric relaxation spectroscopy (DRS). Small-angle neutron and X-ray scattering (SANS/SAXS) probe more extended >1 nm sizes. Computational studies, mainly based on molecular dynamics (MD) simulations and density functional theory (DFT), have provided structural information and insight into the driving forces for cluster and ion pair formation. While the limitations of each of these methods are outside the scope of this work; here we only note that the time scales probed will impact the results.
Nanometric clusters in aqueous ZnSO4 have been hypothesized based on self-diffusion and conductance measurements showing positive cation–cation and anion–anion correlations, even at low concentrations (0.05 M).23 More recently, an experimental study using DRS revealed the presence of CIPs and SSIP,24 with both free ions and ion pairs being strongly hydrated (>1 hydration shells).24 The mean activity coefficient of 1 M ZnSO4 was reported to be 0.042 (where the mean activity coefficient for each component in an ideal solution = 1), demonstrating deviations from ideal solution behavior, due to significant molecular interactions.25 There are several studies suggesting the presence of nanometric clusters for sulfate anions and other divalent metals including magnesium26–30 and calcium.31 However, further studies are required to understand the extent of nanocluster formation, the nature of nanoclusters (including their composition, size, polydispersity, charge, geometry and lifetimes) and their impact on ion transport within these electrolytes.
In this study, as summarized in Fig. 1, we synergistically integrate experimental measurements with computational modeling to unravel the nature of nanoclusters in aqueous ZnSO4 electrolytes. Utilizing SAXS and XPDF analyses alongside MD calculations and forward scattering simulations, we achieve detailed characterization of the size, structure, composition, and distribution of nanoclusters. We then determine the transport from MD and compare it to diffusivity and conductivity measurements, using pulse-field gradient nuclear magnetic resonance spectroscopy (PFG-NMR) and electrochemical impedance spectroscopy (EIS), respectively. We relate the transport measurements to the cluster size distributions determined by X-ray scattering techniques. Our findings reveal a highly polydisperse system of transient nanoclusters composed of hydrated zinc ions and sulfate ions that reduces the ionic conductivity. This combined approach provides a novel method of interrogating electrolyte properties and yields valuable insights for the rational design of electrolytes for sustainable energy storage and other complex liquid systems.
Herein, we use the term nanocluster to describe the measured microheterogeneities in electron density observed by SAXS. These microheterogeneities represent deviations from ideal solution behavior and have historically been referred to as ‘ion pairing’. These microheterogeneities encompass various forms of ion associations, including SSIPs, where cations and anions are connected by solvent molecules; CIPs, where cations and anions are directly connected without intervening solvent; and larger ion clusters (also referred to as nanometric aggregates) that form more complicated structures involving cations, anions, and solvent molecules. As discussed below, we find clusters to be composed of sulfate ions (SO42−) and zinc ions hydrated by 6 water molecules (Zn(H2O)6); a coordination number of six is widely accepted in the literature.26,34–38 For clarity, here we define nanocluster ‘size’ as the ion count which is the number of ions within a nanocluster (as opposed to dimension which is dependent on the geometry) with the expression (Zn(H2O)6)x(SO4)y used to describe a cluster containing x hydrated zinc units and y sulfate ions (cluster size = x + y ions). For simplicity, we do not include sulfate's flexible hydration shell.39–46
To understand the local structure of these nanoclusters, we performed XPDF and MD simulations, obtaining the reduced pair distribution function, G(r), which are shown in Fig. 2b and c. We observe good agreement between the experimental G(r) with MD simulations, shown in Fig. 2b for 0.5 M ZnSO4, where the MD simulations capture the main peak positions and relative intensities found in the measurement, validating the short-range correlations from the MD model. Slight differences in the relative intensities and peak positions may be attributed to the MD force field not perfectly representing the interatomic potentials and/or experimental error for intensities. A clearer understanding of the local structural changes with concentration can be obtained from the differential pair distribution function (dG(r)), shown in Fig. 2d, where the G(r) of a lower concentration solution is subtracted from a higher concentration ZnSO4 solution. The dG(r) highlights changes in the local structure with increasing concentration and is valuable for assessing local cluster structure. Fig. 2d shows good agreement between the differential XPDF for the experimental measurements and MD simulations for 0.5–0.1 M. Correlations identified by the MD simulations are given in more detail in Table S1.†
As expected, the correlations associated with sulfate ion and hydrated zinc increase with increased salt concentration (Fig. 2c and d); specifically the intensity of the correlations associated with the sulfur–oxygen bond within the sulfate ion at approximately 1.54 Å (S–O(SO4)) and zinc–oxygen interaction at 2.1 Å (Zn–O(H2O)), where the subscript refers to the molecule/ion where the element is located. The Zn–O(H2O) correlation distance (position) does not change with increasing salt concentration (Fig. 2c); six water molecules were found in the first solvation shell by MD at all concentrations (0.1–1 M) consistent with the literature where zinc is widely accepted to have six water molecules tightly bound in its first solvation shell in an octahedral geometry.26,34–38 With increasing concentration, the peak at 2.8 Å, related to the hydrogen bonding network in pure water (O–O bond), decreases (Fig. 2c).47 Based on MD simulations, the broad but well-defined peak between 4 and 5 Å (Fig. 2b and c) is due to the second solvation shell of Zn2+ cations which has contributions from both H2O and SO42− anions. In Fig. 2d, the dG(r) shows peaks around 4.2, 5 (black arrow) and 5.8 Å. Using the MD to understand these data, we find that the peak at 4.2 Å primarily arises from correlations between oxygen atoms of water molecules solvating Zn2+ (O(H2O)–Zn–O(H2O)). The correlations around 4.2 and 5.0 Å are attributed to Zn–SO4 solvent-separated ion pairs, both as isolated pairs and within larger nanoclusters. The broad correlations around 5.8 Å and subtle correlations above 7 Å are attributed to the ordering of water, perhaps enhanced by the polarizing nature of zinc (Fig. S1 and S2†).36,38 Although Zn–SO4, Zn–water, and SO4–water correlations extend to distances beyond ∼5.5 Å, these nanocluster correlations are relatively weak; MD simulations show that for distances above ∼5.5 Å the G(r) is dominated by water–water interactions (Fig. S1†). Aside from a larger mean S–O bond distance measured by XPDF compared to the MD (Fig. 2b and d), we observed excellent agreement between MD and experiment. Comparison to the SAXS data, which is unchanged with concentration, suggests that more nanoclusters form as opposed to larger nanoclusters at higher concentrations.
The agreement between experimental and simulated G(r) and dG(r) demonstrates that the MD simulations effectively capture the short-range (here up to ∼8 Å) structural features of (Zn(H2O)6)x(SO4)y nanoclusters and water ordering. As seen in Fig. 2, above 8 Å there are no observable correlations in G(r). Building on this very good agreement between the simulations and measurements, we employ this understanding to investigate the longer-range correlations (above 8 Å) probed by SAXS and gain insights that were previously inaccessible.
![]() | ||
Fig. 3 Schematic showing a few examples of (a) 2-ion Zn(H2O)6SO4, (b) 3-ion (Zn(H2O)6)2SO4, (c) 4-ion (Zn(H2O)6)2(SO4)2 clusters. Dark gray spheres represent zinc ions, orange spheres represent sulfur ions, red spheres represent oxygen atoms, and light pink spheres represent hydrogen atoms. Dashed lines show interactions of oxygen on the sulfate with hydrogens in water molecules in the zinc solvation shell. (d) Statistical analysis of simulated clusters in 1 M ZnSO4 solution using our clustering algorithm based on distance thresholds. The x-axis is the cluster size (number of ions in each cluster) while the y-axis represents the normalized probability of finding a Zn or SO4 ions in each cluster size. This normalization is calculated by dividing the number of ions in clusters of a specific size by the total number of ions in all clusters, thereby reflecting the relative abundance of each cluster size. Gray, orange, and purple bars indicate the fraction of ions that belong to clusters rich in Zn2+, rich in SO42−, and charge-balanced, respectively. Another representation of these data is given in Fig. S9.† (e) Small-angle X-ray scattering data of 1 M ZnSO4 in water compared with the forward simulated SAXS profile calculated from the sum of all nanoclusters from the MD simulations (d) with a scaling factor accounted for the total number of scatters in the probed volume. This poor agreement motivated the method described below. |
To gain a quantitative understanding of ZnSO4 nanoclusters, we developed a clustering algorithm that identifies ion nanoclusters based only on distance thresholds within each MD snapshot. To ensure that our statistical analysis is representative, we analyzed 1400 snapshots over a 14 ns trajectory from the 1 M ZnSO4 MD simulation (each containing 70 Zn and 70 SO4 ions). The size distribution and composition of these nanoclusters determined by this algorithm are shown in Fig. 3d. The analysis reveals a highly polydisperse distribution of nanoclusters from isolated ions up to 22-ion nanoclusters. In Fig. 3d, the number fraction of ions with a particular cluster size (isolated ions are defined as cluster size of 1), with the ratio of ions that belong to Zn-rich, SO4-rich and balanced nanoclusters shown by the colors. The average charge per cluster size are shown in Fig. S8,† assuming a formal charge of +2e for the Zn and −2e for SO4 ions. There is a higher number of free Zn ions compared to SO4 ions (Fig. 3d). Even-numbered clusters generally exhibit a balanced ratio of Zn to SO4 ions, tending towards charge neutrality. Odd-numbered clusters and larger clusters tend to be SO4-rich with an increasingly average negative charge. Sulfate-rich nanoclusters are anticipated from the literature due to sulfate's flexible solvation shell and its ability to form clusters through different mechanisms, such as the equilibrium between bisulfate and sulfate.39–46
The clustering algorithm begins by selecting a zinc ion or the central sulfur atom of a sulfate ion as the starting point and identifying all neighboring zinc and sulfur within the specified distance thresholds (vide infra). These ions are grouped to form the initial cluster. For each ion added to the cluster, the algorithm searches for the newly added ion's neighbors within the distance threshold and incorporates these into the cluster. This process is repeated iteratively for each newly added ion, expanding the cluster to include all connected ions, and the algorithm terminates when no additional neighbors are within the distance thresholds. After defining the cluster with the zinc ions and sulfur atoms from SO42− ions, solvating water molecules surrounding the zinc ions and the oxygen atoms bonded to the sulfur atoms are incorporated into the cluster. The distance thresholds are Zn–S (6.4 Å), S–S (7.5 Å), Zn–Zn (6.6 Å), Zn–O(H2O) (3.1 Å) and S–O (1.51 Å) and were chosen as the first minimum between the first two peaks of their partial pair distribution functions of the MD simulations (Fig. S3–S7†).48 The use of the first solvation shell as the distance threshold is rooted in capturing the most significant and relevant atomic correlations without overestimating cluster connectivity, which might lead to unrealistic cluster sizes. Importantly, variations (±0.3 Å) in these thresholds do not significantly affect the results, demonstrating the robustness of the algorithm. Note that for Zn–S the first peak (CIPs) is not used for distance threshold as the coordination number (0.01) is very small, indicative of negligible CIPs. Using the first solvation shell as the distance threshold ensures that all correlations within the range of significant atomic proximity are included, even those that are transient or less enthalpically favorable. We note this algorithm includes transient clusters formed either through electrostatic or stochastic fluctuations. Despite repulsive forces, the probability of finding a zinc-to-zinc or sulfate-to-sulfate correlation within the defined distance threshold is non-zero, highlighting the role of entropy in forming short-lived ion clusters.
While our work concludes that the majority of ions are in clusters, previous studies using DRS found the majority of ions were free; a comparison is provided in Table S2.† Interestingly, these authors also reported CIPs, however, they stated that the CIP fraction was an unresolved mixture of CIPs and dynamically-retarded (‘slow’) water.24 The difference between our conclusions and those of the DRS likely arise because of differences in time scales probed. Dielectric spectroscopy probes ion dynamics in the 0.1–100 GHz regime, corresponding to ∼10 ps to 10 ns timescales, while SAXS and XPDF probes near-instantaneous (fs) structure, time- and space-averaged over the measurement time and probed volume. The nanoclusters observed with SAXS and characterized in our MD simulations have lifetimes much shorter than ns, therefore DRS is insensitive to the nanoclusters described here. Furthermore, even if these nanoclusters are long-lived, they could rotate too slowly and/or have a net dipole that is too small to be detected by DRS.
As the experimental XPDF results (Fig. 2) align well with the MD results, we are confident of the reliability of the nanocluster's short-range structure. Therefore, the MD nanocluster configurations provide an exciting opportunity to interpret the SAXS data without the need of a model requiring assumptions regarding the shape, polydispersity or atomic positions of atoms within nanoclusters. To better understand how clusters of each size contribute to the SAXS data, we created average scattering curves for nanoclusters of the same ion count. We note that a wide range of nanocluster sizes are present, which could be defined by a variety of parameters such as correlation length, volume, geometry or radius of gyration; therefore, we define nanoclusters by their ion count for simplicity. Note, we report the size range (electron weighted radius of gyration) for different cluster ion counts based on the atomic positions determined by the MD in Table S3.†
Here we first forward-simulate the SAXS for each unique nanocluster configuration obtained from the MD simulation from the nanocluster's specific arrangement of atoms. Then, these individual SAXS profiles were averaged for nanoclusters with the same ion count to generate an ensemble SAXS pattern for that ion count (Fig. S10†). Examples of the calculated SAXS profiles are given in Fig. 4, where the intensity of these simulations has been scaled to the experimental SAXS data. This qualitative analysis reveals that smaller nanoclusters, specifically those with three ions, model the high-q region (corresponding to small cluster sizes) reasonably well but inadequately model the low-q region (corresponding to larger cluster sizes) of the experimental SAXS profile. Similarly, larger nanoclusters agree with the low-q region but do not adequately represent the high-q region. Nanoclusters containing four ions exhibit excellent agreement with the experimental data across a broader q range, with some deviation at higher q. Whilst nanoclusters with an ion count of 4 can adequately fit the experimental data, this overlooks the broader complexity inherent in the system. This simplification can lead to incomplete conclusions about the nature of the interactions and structures present, which in turn can inhibit our ability to understand and thus predict behavior.
![]() | ||
Fig. 4 Average SAXS forward simulations from the SAXS profiles of each cluster configuration identified in the MD containing (a) 3, (b) 4, (c) 5, (d) 6, (e) 7, and (f) 8 ions. The pink markers represent the experimental SAXS profile for 1 M ZnSO4 solution, the green dashed line represents the forward simulated average SAXS profile for each cluster ion count, the gray dotted line represents the fitted constant background and the black line represents the overall fit (cluster + background). Note that small clusters, such as clusters with ion count of 3 fit the high q region well, and larger clusters with ion count more or equal to 5 fit the low q region well, highlighting that a distribution of cluster sizes is necessary to capture the experimental data. (As q is roughly inversely proportional to the cluster size, the low and high q region correspond to large and small clusters respectively.) The radius of gyration for each cluster ion count as well as the standard deviation is given in Table S3.† |
To best relate the experimental SAXS to the cluster size distribution, we fitted the experimental data using a weighted sum of these average SAXS profiles for each cluster size (Fig. S10†). In SAXS, the intensity of a scattering curve is proportional to the number density of scatterers. Thus, the fitted weight of each SAXS profile is proportional to the number density of nanoclusters of the corresponding size within the experimental ZnSO4 solution. At first, the distribution of weights was constrained in the fit to a predetermined Schulz-Zimm distribution (Fig. S11†). However, a variety of nanocluster distributions may result in good fits. To better explore the parameter space and account for many possible fits, we employed a Monte Carlo Markov Chain (MCMC) method implemented via the Differential Evolution Adaptive Metropolis (DREAM) algorithm. This MCMC method enabled us to sample from a variety of cluster size distributions that may fit the experimental SAXS data.49,50 In this MCMC method, the weight of each cluster size's average scattering profile was allowed to independently vary, and the MD-derived cluster size distribution served as an initial distribution of weights to initialize each Markov chain. This approach produced a well-converged posterior distribution with excellent agreement with the experimental SAXS data, as shown in Fig. 5a. Note that the MCMC forward simulation was done with 120 MD snapshots rather than with 1400 snapshots for efficiency. However, the distribution of the 120 snapshots is representative of the 1400 snapshots (Fig. S12†).
![]() | ||
Fig. 5 (a) Small angle X-ray scattering data of 1 M ZnSO4 in water (shown in pink), overlaid with the fitted SAXS profile (black line). The fit was generated through a weighted sum of averaged simulated SAXS curves for each cluster size. Weights were fitted using a Monte Carlo Markov-chain algorithm (25 chains, 50![]() |
As seen in Fig. 5 and Fig. S11,† we consistently observed a wide distribution of nanoclusters and the absence of large nanoclusters (>10-ion clusters), irrespective of the approach used to fit the SAXS data. In Fig. S11,† we explored different fitting methods, including Levenberg-Marquardt optimization with weights derived from a Schulz-Zimm distribution model,51–53 and a direct fitting approach based on a modified MD cluster size distribution. While these approaches provided excellent fits, the MCMC method demonstrated greater flexibility by accounting for a broader range of possibilities, offering a statistically robust distribution of solutions. Based on this, we consider the MCMC method implemented through DREAM to be a reliable approach for exploring potential fits to SAXS data that may not be uniquely described by a single model such as directly using the MD or a Schulz-Zimm distribution. We encourage the independent validation of our methodology on other systems.
The shaded region in Fig. 5b indicates the posterior distribution of probable fits, illustrating that multiple ion cluster distributions can adequately fit the SAXS data, especially for nanoclusters smaller than 6 ions where the possibilities/error bars are relatively wide. Despite variations between the MD and the resulting distribution of fitted weights from the DREAM analysis (Fig. 5b and Fig. S13†), the distribution determined by MD generally falls within the posterior distribution for nanocluster sizes up to 6 (Fig. 5b). While MD simulations predict the presence of large nanoclusters, the experimental SAXS data (low q region) show that these nanoclusters are unlikely to form in significant quantities, shown in Fig. 5b. Meanwhile, nanoclusters with ion counts of 3 and 4 are the most common nanocluster size, as evidenced by their larger probabilities in the posterior distribution (Fig. 5b), suggesting that these configurations may be energetically favorable. Further investigation into the thermodynamics of nanocluster formation is needed to fully understand the preferential formation of certain nanocluster sizes.
In addition, we forward-simulated the G(r) for both the MCMC and MD cluster size distributions. As shown in Fig. S14,† G(r) is not sensitive to differences in these two different cluster size distributions. This insensitivity explains why determining cluster size distributions solely based on MD and experimental G(r) is challenging. The variety of size distributions that adequately fit the experimental data highlights the inherent complexity in modeling solutions with polydisperse nanoclusters—dynamic arrangements exhibiting a wide range of sizes, geometries, and compositions. This polydispersity arises from thermodynamic and kinetic factors driving the formation of diverse nanocluster populations, emphasizing the inadequacy of simplistic models that assume uniformity in nanocluster characteristics. Importantly, based in the MD these nanoclusters are short-lived, existing transiently on ∼picosecond timescales, further underscoring the dynamic nature of these systems.
![]() | (1) |
![]() | (2) |
![]() | (3) |
Based on MD simulations, the diffusivity was found to be 6.4 × 10−10 and 9.1 × 10−10 m2 s−1 for Zn and SO4 ions, respectively. The system size correction was applied using the Yeh and Hummer method55,56 with the calculated viscosity. Using these values, sNE was found to be 23.2 S m−1. When the Einstein relation is used, which includes ion correlations, the sE was found to be 18.7 S m−1, lower than sNE by about 20%. These values are higher than the experimental conductivity of 4.42 S m−1, obtained using EIS (Fig. S15†). The experimental conductivity is consistent with previously reported values in the literature.57,58 This discrepancy between MD and experimental results is similar to our prior studies, where the diffusivity (probed experimentally over smaller length scales) shows good agreement but the conductivity (probed experimentally over larger length scales) was found to be overestimated in simulations.59 Further studies to validate the force field with experimental transport measurements, such as viscosity and NMR diffusion studies, will shed further light on whether this discrepancy is a physical effect, such as solvation drag effects. The ratio sE/sNE is called inverse Haven ratio (or ‘degree of uncorrelated ion motion’), which is a measure of the ionicity of the electrolyte. For an ideal electrolyte without any ion correlation the inverse Haven ratio equals one, while for the 1 M ZnSO4 solution the inverse Haven number was found to be 0.81, indicating that ion correlations decrease conductivity but not dramatically. Furthermore, the apparent transference numbers were determined from the MD to be 0.41 for Zn2+ and 0.59 for SO42−, reflecting their relative contributions to ionic transport in the electrolyte.
Experimentally, the diffusivity of Zn2+ measured using PFG-NMR was slightly lower, at 4.24 × 10−10 m2 s−1, compared to the MD-calculated value (6.4 × 10−10 m2 s−1). This difference may be due to more clusters or longer-lived transient clusters in the real system, therefore restricting ion mobility. This interpretation could be connected to higher probability of smaller clusters observed experimentally via SAXS compared to calculated in the MD simulations (Fig. 3e). The difference in diffusivity between experiment and simulations, combined with the difference between calculated sNE and sE, suggests that the ions are weakly correlated, and the clusters are relatively short-lived. Rather than a purely hopping mechanism, ion transport in zinc sulfate electrolytes involves a combination of vehicular and hopping motion. While zinc ions move with their hydration shell (first solvation shell containing 6 water molecules) via vehicular motion, the interactions between zinc and it's second solvation shell are transient, similar to the hopping mechanism proposed for sulfate ions. This interplay allows for efficient ion migration, where ions dynamically migrate between transient clusters, preserving high conductivity.
As shown in our previous studies, sE can be divided into contributions from different correlations.60,61 The same analyses were carried out in the current study and the results are shown in Table 1. As indicated by the Einstein relation and illustrated in Fig. 6a to aid interpretation, for ions with like charges, an overall negative or positive contribution to the Einstein relation corresponds to the motions of these ions being anticorrelated or correlated, respectively. Here we find a negative contribution for the zinc ions (Zn–Zn); this may be due to coulombic repulsion of like ions. Meanwhile we hypothesize that the small positive contribution to the Einstein relation observed for the sulfate ions (SO4–SO4) may be due to the lower coulombic repulsion between sulfate ions due to their delocalized charge and therefore weaker interactions and less specific correlations and/or sulfate's ability to form transient clusters with protons. We observe a negative contribution to the Einstein relation for ions with opposing charges (Zn–SO4).
![]() | ||
Fig. 6 (a) Illustration to aid the interpretation of negative or positive contributions to the ionic conductivity determined by the Einstein relation (sE) given in Table 1, which depends on the ion's charge (where stripped and clear are used to represent different charges) and direction of travel (denoted by the arrow). This schematic depicts the overall average contribution of ions in the simulation box, not individual atom pairs. (b) Normalized probability of finding isolated or clustered zinc or sulfate ions. The normalization is calculated by dividing the number of zinc or sulfate ions in each state by the total number of zinc and sulfate ions in the system. (c) Illustration showing the coordination number (CN), residence time (τ) and expected motion calculated from the MD simulations and given in Table 2 for [ion]![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
Contribution to σE (S m−1) | Motion type | |
---|---|---|
Zn | 9.573 | |
SO4 | 13.573 | |
Zn–Zn | −3.227 | Anticorrelated |
Zn–SO4 | −1.673 | Correlated |
SO4–SO4 | 0.490 | Correlated |
As shown in Fig. 3d and 6b, the results of our transport measurements and calculations can be explained by our observation of the formation of nanoclusters which on average tend to be sulfate-rich together with some isolated zinc ions. Overall, the negative contributions to the ionic conductivity suggest that the Zn–Zn and Zn–SO4 correlations reduce the number of free, independently moving charge carriers, thereby lowering the overall ionic conductivity.
The residence times of ions/molecules in the first and second solvation shell of zinc and sulfate ions provide insights into overall transport processes. The results are summarized in Table 2 and illustrated in Fig. 6c. Consistent with our model of hydrated zinc and sulfate units forming nanoclusters, we observe long residence times for zinc's first solvation shell containing water as well as zinc's second solvation shell containing sulfate ions. For water molecules in zinc's first solvation shell, the residence time correlation function almost does not decay during the simulation time scale, indicative of a strong and stable hydration shell around Zn2+ containing 6 water molecules; this structure has been widely reported.26,34–38 Furthermore, the 2nd shell Zn–SO4 solvation shows significant long lifetimes (∼0.2 ns) with a coordination number of 2.4. Note, whilst the Zn–O(SO4) interaction in the first solvation shell also shows a relatively long lifetime (∼1 ns) compared to other interactions, the coordination number is very small (0.01).
Solvation shell | Cutoff (Å) | CN | τRT (ns) | Travel distance (Å) | Motion type | |
---|---|---|---|---|---|---|
Zn![]() ![]() |
1st | 3.1 | 0.01 | 0.989 | 17.74 | Vehicular |
2nd | 5.5 | 2.39 | 0.212 | 8.22 | Hopping | |
Zn![]() ![]() |
1st | 3.1 | 5.99 | ≫14 | ≫3.1 | Vehicular |
2nd | 5.2 | 15.50 | 0.041 | 3.62 | Hopping | |
O(SO4)![]() ![]() |
1st | 2.7 | 3.79 | 0.007 | 1.84 | Hopping |
2nd | 4.7 | 24.61 | 0.034 | 4.01 | Hopping |
Using the measured diffusivity and the calculated residence times, the distance an ion can travel during the period of the residence time was calculated and included in Table 2. It is clear that Zn2+ and its first hydration shell can travel a distance much longer than the cluster size indicated by the corresponding cutoff distance before the cluster breaks, suggesting a vehicular transport mechanism, where the ion and its solvation shell move together through the electrolyte.60,62 On the contrary, other interactions are more likely to transport through a ‘hopping’ mechanism, involving frequent breaking and reforming of interactions. We note that the SAXS reveals these nanoclusters are present but provides no insight into their lifetimes.
The presence of some of the ions within nanoclusters (Fig. 6b) leads to a 20% reduction in Einstein conductivity compared to the Nernst-Einstein conductivity, indicating that extensive and varied nanoclusters (Fig. 5) modestly hinder ion transport. Their transient nature prevents a significant decline in overall conductivity. These findings further reinforce the role of transient ion clusters in mediating ionic transport. While the strong zinc–water interaction stabilizes the ion's local environment, the dynamic exchange of both hydrated zinc and sulfates between solvation shells contributes to the continuous evolution of clusters and enables efficient ion transport.
Overall, this analysis reveals that the complicated clustering structure of the ZnSO4 electrolyte introduces significant heterogeneity into the dynamics and transport behavior. While transient ion clusters reduce diffusivity and ionic conductivity compared to ideal systems, their short-lived and dynamic nature ensures that ions retain significant mobility, enabling effective charge transport.
We envision that the approach established here will serve as a foundation for future studies to experimentally investigate nanocluster properties, such as dynamics or ion pairing. By using our methodology to interrogate how the nanocluster's properties vary with conditions (such as concentration, temperature or pH), and strategies to improve performance, such as the use of additives, we can meaningfully understand how the presence of nanoclusters affects battery performance, such as cycling stability and overpotentials. In addition, our methodology is applicable to other systems where nanometric aggregates are formed, for example other electrolytes or perovskite solutions. A deeper understanding of the nature of nanoclusters in electrolytes will facilitate more comprehensive studies on ion transport, electronic structures, solvation processes and interfacial mechanisms; all of which are essential for the rational design of superior electrolytes for a wide range of applications.
![]() | (4) |
![]() | (5) |
To calculate the reduced pair distribution function G(r), we do the sine Fourier transform of the reduced structure function:
![]() | (6) |
Here qmax is the maximum q value and we used 30 Å−1, and L(q) is the Lorch function to account for termination effects:
![]() | (7) |
![]() | (8) |
Three salt concentrations, 0.1 M, 0.5 M and 1 M, respectively, were studied. The box compositions are shown in Table 3. Periodic boundary conditions were applied in all three dimensions. The long-range electrostatic interactions, with a real space cutoff of 12 Å, were calculated based on the particle–particle particle–mesh (PPPM) method.76 The same cutoff was used for van der Waals interactions and a tail correction was applied.77 The systems were first simulated for 2 ns in the isothermal-isobaric (NPT) ensemble at 1 atm pressure (isotropic volume fluctuations) and 298 K. The equilibrated simulation boxes were simulated for 14 ns as the production simulation in the canonical (NVT) ensemble at the same temperature. The Nose–Hoover thermostat78 and the extended Lagrangian approach79 were applied to control the temperature and pressure, respectively. Five independent trajectories were generated using different initial atomic velocities. The time step for all the simulation was 1 fs and snapshots were saved every 1000 steps.
ZnSO4 concentration (M) | Number of ZnSO4 | Number of water |
---|---|---|
0.1 | 7 | 3866 |
0.5 | 30 | 3320 |
1.0 | 70 | 3866 |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5eb00119f |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |