Significance
In electrolytes, mobile ions surround charged colloids, screening their electric fields. Such electrolyte–macroion interactions modulate the structure, function, and assembly of biological macromolecules. Classical theories predict increasingly screened electrostatic interactions with elevated salinity, enabling more compact arrangements of like-charged components. Here, X-ray scattering measurements and molecular dynamics simulations demonstrate that this principle breaks down at high salt concentrations. Assemblies of DNA-coated nanoparticles switch from contraction with added salt at low salinities, to expansion at high salinities. Interactions between electrolyte ions drive this contraction-to-expansion transition. These results demonstrate the significance of ionic correlations in concentrated electrolytes and should prove useful in the development of theories and electrolyte-based technologies.
Keywords: electrolyte, underscreening, DNA, self-assembly, SAXS
Abstract
Electrostatic forces in solutions are highly relevant to a variety of fields, ranging from electrochemical energy storage to biology. However, their manifestation in concentrated electrolytes is not fully understood, as exemplified by counterintuitive observations of colloidal stability and long-ranged repulsions in molten salts. Highly charged biomolecules, such as DNA, respond sensitively to ions in dilute solutions. Here, we use non-base-pairing DNA-coated nanoparticles (DNA-NP) to analyze electrostatic interactions in concentrated salt solutions. Despite their negative charge, these conjugates form colloidal crystals in solutions of sufficient divalent cation concentration. We utilize small-angle X-ray scattering (SAXS) to study such DNA-NP assemblies across the full accessible concentration ranges of aqueous CaCl2, MgCl2, and SrCl2 solutions. SAXS shows that the crystallinity and phases of the assembled structures vary with cation type. For all tested salts, the aggregates contract with added ions at low salinities and then begin expanding above a cation-dependent threshold salt concentration. Wide-angle X-ray scattering (WAXS) reveals enhanced positional correlations between ions in the solution at high salt concentrations. Complementary molecular dynamics simulations show that these ion–ion interactions reduce the favorability of dense ion configurations within the DNA brushes below that of the bulk solution. Measurements in solutions with lowered permittivity demonstrate a simultaneous increase in ion coupling and decrease in the concentration at which aggregate expansion begins, thus confirming the connection between these phenomena. Our work demonstrates that interactions between charged objects continue to evolve considerably into the high-concentration regime, where classical theories project electrostatics to be of negligible consequence.
Concentrated electrolytes are relevant to a wide variety of fields, ranging from the development of electrochemical devices, such as supercapacitors (1), to extremophile biology (2). Electrostatics in dilute solutions of salts have been extensively described by mean-field theories such as the Debye–Hückel theory (3). In such theories, mobile salt ions reorganize in response to the electrostatic potential of the charged macroions (e.g., colloids and membranes). In particular, salt ions concentrate near oppositely charged surfaces and screen their electrostatic field. The effective range of electrostatic interactions (screening length) is expected to decay with salt concentration C as Therefore, a direct extension of this theory to high salinity would imply that the range of charge-driven effects decreases to negligible distances. In reality, interactions between the dissolved ions, which are ignored in the Debye–Hückel theory, become increasingly important at high salt concentrations, resulting in ionic correlations that are influenced by ion size and valency (4–6) and strongly regulate ion distributions in confinement by dielectric surfaces (7). These ionic correlations in concentrated monovalent and multivalent electrolytes drive overscreening of charged surfaces (8) and stabilize colloidal nanoparticles against aggregation in molten salts (9). Furthermore, high degrees of ionic organization are theorized to underlie the unexpectedly long electrostatic screening lengths measured via surface force apparatus in highly saline solutions (10–12). Interest in the generality of such findings has motivated additional research (13–19), which has revealed the need for further investigations through experimental studies probing electrostatic interactions in concentrated salt solutions.
In the present study, we investigate the assembly of highly negatively charged DNA-functionalized Au nanoparticles (DNA-NP) in concentrated aqueous solutions of alkaline earth chlorides: CaCl2, MgCl2, and SrCl2. The grafted DNA sequences are designed to exclude interparticle attractions via DNA base-pairing. Even then, these dissolved salts at moderate concentrations (≲1 M) induce attractive interactions between the like-charged particles, driving their assembly into crystalline and noncrystalline aggregates (13, 20). We utilize a combination of small-and-wide-angle X-ray scattering (SAXS/WAXS) to analyze the structures formed by these nanoparticles over the whole accessible range of electrolyte concentrations (via SAXS), while also assessing the ion–ion correlations present in the bulk electrolyte (via WAXS). As would be expected from Debye–Hückel theory, SAXS measurements show that the observed structures contract with added salt. However, this trend holds only until a threshold concentration, above which the structures expand as salt concentration increases. This lattice reexpansion is connected to enhanced cation–anion ordering in the bulk solutions resolved by WAXS measurements and is explained through molecular dynamics (MD) simulations, which reproduce the nonmonotonic swelling of the nanoparticle aggregates. Overall, the combined experimental and computational results demonstrate how interactions mediated by dissolved ions continue to evolve at high salt concentrations, and provide insight into how ion–ion correlations manifest in interactions between charged nanoscale objects dispersed in concentrated electrolytes.
The nanoparticles utilized in this study were prepared by grafting thiolated single-stranded DNA molecules consisting of 35 thymine bases (T35) onto Au cores (nominal radius, RAu = 4.5 nm). The grafting density was ~ 60 to 80 DNA/Au nanoparticle, with loading varying between synthesis batches (SI Appendix, Table S1). Unless specified otherwise, samples for measurements were prepared with DNA-NP concentration of 50 nM, corresponding to a dilute suspension (volume fraction ~ 10−3). Additional sample preparation details are provided in SI Appendix. Despite the exclusion of canonical base pairing between neighboring DNA-NP, such particles assemble into colloidal crystals in solutions containing sufficient amounts of divalent cations. In such a condensed state, the proximity of the AuNP cores causes redshifting of the absorbance corresponding to the surface plasmon resonance, resulting in the suspensions changing color from red to purple (21) (Fig. 1A). Given sufficient time, the nanoparticle aggregates precipitate from the solution. Lowering the salt concentration results in the “melting” of the aggregates and redispersion of the particles, as visualized by the restoration of the suspension’s red color (Fig. 1A). This reversibility implies that the assemblies are equilibrium phases, with salt-concentration-determined structures.
Fig. 1.
Assembly of DNA-NPs in CaCl2 solutions. (A) Schematic (Top) and photographs (Bottom) of CaCl2-induced reversible crystallization of DNA-functionalized nanoparticles (DNA-NP). Increasing the salt concentration results in nanoparticle aggregation, which is reflected in the suspension color changing from red to purple (Middle). Decreasing the salt concentration restores the red color, showing that the aggregation of the nanoparticles is reversibly controlled by salt concentration. Note that while the salt concentration is varied, the DNA-NP concentration is held fixed at 50 nM. (B) DNA-NP structural phase diagram and center-to-center interparticle separations, Dnn, as a function of CaCl2 concentration. Note that the effective volume per particle, calculated after correcting for the packing fractions in different assembly structures, shows the same trend as Dnn (SI Appendix, Fig. S6). (C and D) Representative structure factors with corresponding model fits for DNA-NP face-centered cubic [FCC, (C)] and body-centered cubic [BCC, (D)] lattices. (E) Structure factors for the assemblies formed at high CaCl2 concentration, which display no long-range order. The position of the primary diffraction peak initially shifts to higher q with added CaCl2, consistent with a contracting structure; this trend reverses at high CaCl2 concentrations (>3.3 M). In C−E, the structure factors are offset vertically for clarity.
To investigate the assembly structures, we analyzed DNA-NP dispersions over the whole accessible range of CaCl2 concentrations (0 to 5.3 M) via in situ SAXS. All dispersions were allowed to equilibrate in epoxy-sealed quartz capillaries for 24 to 36 h prior to measurements. For details of SAXS measurements and data processing, see SI Appendix. Fig. 1 C−E, show examples of SAXS-derived structure factors S(q) at varied salt concentrations. Here, q = 4π sin θ/λ is the scattering vector magnitude, 2θ is the scattering angle, and λ is the X-ray wavelength (here, λ = 0.729 Å). S(q) results from the interference of X-rays scattered from distinct particles in the assemblies and yields assembly phases and interparticle distances. DNA-NP dispersed in [CaCl2] ≲0.25 M remain suspended, with the suspension retaining its red color. However, very weak diffraction peaks are discernible in S(q) for [CaCl2] ≳ 0.2 M (SI Appendix, Fig. S1), likely corresponding to fluid-like correlations between small numbers of particles. Such dispersed phases are referred to as precrystalline. By contrast, Fig. 1 C and D show sharp diffraction peaks corresponding to DNA-NP organization into crystal lattices over [CaCl2] ~ 0.3 to 1.7 M. For [CaCl2] ≳ 1.7 M, much broader intensity modulations corresponding to short-ranged positional correlations between DNA-NP are observed (Fig. 1E). Detailed analysis of these structure factors reveals that the DNA-NP assemble into face-centered cubic lattices (FCC) for [CaCl2] ~ 0.3 to 0.8 M (Fig. 1C), as observed in our previous work (20). Furthermore, here we find that for [CaCl2] ~ 0.8 to 1.7 M, DNA-NP assemblies convert from FCC to body-centered cubic (BCC) (Fig. 1D). Above [CaCl2] ~ 1.7 M, the intensity profiles exhibit broad peaks with a ratio between the positions of the first 2 intensity maxima. This is consistent with random close packing (RCP) (22, 23). These results were reproduced using multiple batches of DNA-NPs (SI Appendix, Fig. S2) and were influenced minimally by decreasing the DNA-NP concentration from 50 to 20 nM. (SI Appendix, Fig. S3).
Apart from the salt-concentration-induced structural transformations, analysis of S(q) (Fig. 1 C−E) also demonstrates that the distances between the nanoparticles vary with CaCl2 concentration. We note that for both FCC and BCC structures, the nearest neighbor distance (Dnn) is related to the position of the principal diffraction peak (q1) as (SI Appendix, Table S2). This relationship also extends to RCP, which lacks long-range ordering, as demonstrated by nearly identical values extracted through this formulation and via Fourier analysis of the S(q) (SI Appendix, Fig. S4 and Table S3). Fig. 1 C−E show that q1 increases (Dnn decreases) with added salt up to [CaCl2] ~ 3.3 M. Over this range, Dnn decreases from ~ 24 to ~16 nm. Assuming that there is minimal interdigitation between the DNA shells from neighboring particles, this corresponds to a contraction of DNA shell from 7.5 to 3.5 nm. These lengths are much smaller than the T35 contour length of ~0.65 × 34 ~ 22 nm (24), implying that DNA acts as a flexible oligomer in compact conformations in the ionic conditions used here. Above the threshold of [CaCl2] ~ 3.3 M, the trend in q1 (and thus Dnn) reverses, signifying that the aggregates expand. The phase transformations of DNA-NP assemblies and the variation of the interparticle separation with CaCl2 concentration are summarized in Fig. 1B.
At first glance, the observed structural transformations, particularly the FCC to BCC transition, are surprising. This is because long-ranged, softer electrostatic interactions expected in the low salt concentration regime favor BCC, as observed in the assembly of like-charged colloids interacting via screened Coulomb potential (23). In that work, the assemblies transformed from BCC to FCC with increasing salt concentration. This is also consistent with observations on colloids and nanoparticles with hard cores and soft polymer shells, for which the BCC phase is favored in systems with longer and more flexible shells, which lead to “softer” interparticle interactions (25, 26). We have analyzed the observed FCC to BCC transformation elsewhere (27) and have shown that interparticle attractive interactions coupled with the dehydration of the DNA shell at elevated salt concentrations lead to the observed FCC to BCC transitions. In this work, we will focus on the surprising contraction and reexpansion of the assemblies at high salinity.
The observed lattice contraction is consistent with a picture where increasing the divalent cation concentration (1) increases attractive interactions between the DNA in the overlap region of the neighboring DNA-NP and (2) screens the repulsions between charged phosphate groups along the DNA strands, allowing more compact DNA conformations. Similar contraction of nanoparticle lattices in a moderate salt concentration regime (<2 M) has also been observed for DNA-NP assembled via Watson–Crick hybridization (28, 29). Therefore, the reexpansion of aggregates above a threshold concentration is nonintuitive. To test whether aggregate reexpansion is an effect of specific Ca2+–DNA interactions or a more general phenomenon, we repeated SAXS measurements in MgCl2 (0 to 4.5 M) and SrCl2 solutions (0 to 3 M), as shown in Fig. 2.
Fig. 2.

Assembly of DNA-NPs in SrCl2 and MgCl2 solutions. (A) Structure factors for DNA-NP assemblies at varied SrCl2 concentrations. The dotted line is a guide to the eye and indicates the variation in the position of the principal peak. The corresponding inter-DNA-NP distances as a function of [SrCl2] are shown in (B). (C and D) Structure factors observed at various MgCl2 concentrations. Note that the position of the principal peak shifts to higher q up to 2 M (lattice contracts with added MgCl2). Above this concentration, added salt leads to lattice expansion. (E) Summary of structural phases and interparticle separations observed as a function of MgCl2 concentration. No detectable scattering peaks were observed in solutions below ~1 M. In A, C, and D, the structure factors are offset vertically for clarity.
There are clear distinctions between the assembly structures observed in CaCl2 (Fig. 1B), SrCl2 (Fig. 2B), and MgCl2 (Fig. 2E). For example, DNA-NP dispersed in SrCl2 solutions do not form precipitates. The suspensions at all concentrations retain their red color. The SAXS-derived structure factors for these dispersions (Fig. 2A) display a weak, broad diffraction peak, indicative of a fluid-like ordering (30), here defined as precrystalline. In MgCl2 solutions, DNA-NP assemble into FCC for [MgCl2] ~ 1 to 2.5 M (Fig. 2C) and convert to structures with short-ranged ordering above 2.5 M (Fig. 2D). This behavior is comparable to that observed in CaCl2 solutions (Fig. 1B), with the exceptions that 1) no BCC phase is observed in MgCl2 solutions and 2) the salt concentration for DNA-NP crystallization into FCC is higher (1 M for MgCl2 vs. 0.3 M for CaCl2). These contrasting properties of the assemblies reflect specific cation–DNA interactions. For example, the observed assembly behavior is consistent with the established understanding that the smaller Ca2+ or Mg2+ cations associate more strongly with DNA than larger Sr2+ or Ba2+ cations (31). Interestingly, the smaller size of Mg2+ (as compared to Ca2+) would suggest a stronger electrostatic interaction with DNA. However, we observe that higher concentrations of MgCl2 are required to induce nanoparticle crystallization and that assemblies in MgCl2 solutions feature larger interparticle separations than those in CaCl2 solutions (Figs. 1B and 2E). We speculate that this weaker interaction arises due to the tendency of Mg2+ to retain its primary hydration shell when interacting with the DNA phosphate backbone (32, 33). Despite these distinctions in the assembly behavior, the SAXS-extracted Dnn values show the same qualitative trend of aggregate contraction followed by reexpansion for assemblies in each of the three salt solutions (Figs. 1B and 2 B and E).
There are some interesting phenomenological observations associated with the observed reexpansion. 1) The reexpansion occurs regardless of whether the structures exhibit crystalline (MgCl2), short-ranged (CaCl2), or liquid-like order (SrCl2). 2) The value of the threshold concentration Cth for the upturn in Dnn is cation-dependent, increasing in the sequence Sr2+ (1.5 M) < Mg2+ (2.0 M) < Ca2+ (3.3 M). A potentially related phenomenon to this reexpansion has been observed in high charge density polyelectrolytes, which precipitate from solution at low concentrations of cationic multivalent ions and polyamines with Z ≥ 3 and redisperse at higher salt concentrations (34, 35). The redissolution was speculated to be due to a charge reversal (36), in which the charge due to adsorbed counterions, at the macro-ion surface, exceeds the charge required to neutralize the macro-ion. However, DNA charge reversal has not been documented in aqueous concentrated electrolytes with Z ≤ 3 cationic proteins or ions, (37). Additionally, this charge reversal model does not consider the association of cations to anions in the bulk (38), which can lead to redissolution without charge reversal (39). Furthermore, if a charge-reversal mechanism were to underlie the presently observed DNA-NP reexpansion, then the Cth would be expected to be lowest for the cation that interacts most strongly with the DNA-NP (Ca2+). The observed Cth sequence is, in fact, the reverse of this prediction. Rather, the sequence follows that of the solubility limits of the corresponding chloride salts in water (40). Notably, for each salt, Cth is consistently ~ 40 to 60% of the salt’s solubility limit in water at room temperature (SI Appendix, Fig. S5). Furthermore, our molecular simulations (discussed below) reveal that the effective electric field outside the DNA shells remains weakly negative at all salt concentrations (SI Appendix, Fig. S10). These observations suggest that it is the interactions between electrolyte ions in close proximity, which become prominent at high salinities, that drive the nonmonotonic interparticle separation behavior.
To gain insights into the electrolyte ion–DNA-NP and interelectrolyte interactions, we performed MD simulations using a coarse-grained model to analyze a dense phase of DNA-NP in equilibrium with MX2 reservoirs (Fig. 3 A and B). The model used implicit solvent, but explicit ions. Due to computational constraints, the simulations utilized smaller nanoparticles, (RAu, MD = 2.25 nm) coated by 18 DNA, each represented by 16 charged (−e) beads with monovalent counterions, here labeled M+ (for more details, see SI Appendix). MD results of the equilibrium ion concentration distributions ([M2+], [X−], [M+]) and Dnn as a function of reservoir concentration are shown in Fig. 3 C and E, respectively. Fig. 3D. illustrates the method for extracting Dnn at a given concentration. Fig. 3E corroborates the experimentally observed (Fig. 1B) Dnn nonmonotonic behavior. The transition molar concentration Cth, MD ~ 2.0 M for the upturn in Dnn is lower than the experimentally observed Cth for Ca2+(~3.3 M) but close to Cth for Mg2+. We note that the implicit solvent utilized in the MD simulations does not consider ion hydration effects. Furthermore, below 2.0 M, in the regime where Dnn decreases with increasing salt concentration, MD simulations (Fig. 3F) reveal that the divalent ions’ molar concentration in the dense DNA-NP phase is higher than in the reservoir Interestingly, in the expansion regime, the concentration of the divalent cations in the DNA-NP phase is slightly lower than in the reservoir . It should be noted that, across the whole salt concentration regime for DNA-NP assembly, the DNA brush contains high concentrations of both cations and anions. Details of the radial ion distribution profiles are provided in the SI Appendix, Fig. S9. In the aggregate contraction regime, these results suggest that the system’s equilibrium is driven by the internal energy decrease due to DNA-M2+ interactions. In contrast, in the aggregate expansion regime, the system’s equilibrium is dictated by 1) the entropy gained through DNA adapting more extended conformations, enabling ions redistribution in a larger volume of the DNA shell, and 2) favorable electrostatic interactions between salt ions.
Fig. 3.

MD simulations on DNA-NP assemblies at different salt concentrations. (A) Coarse-grained model for a DNA-NP. The 16 DNA beads, representing a single DNA, bear a negative elementary charge (−e) and are connected using a harmonic potential. (B) To form the dense phase, 108 nanoparticles are initially arranged in an FCC lattice and surrounded by monovalent counterions. An ion solution containing divalent cations and monovalent anions is in contact with the dense phase. (C) The ionic concentration profiles in the z-direction, after equilibration, for divalent cations, and monovalent cations and anions. (D) Radial distribution functions gNP(r) for DNA-NP are calculated using the equilibrium configurations. (E) The nearest-neighbor separation distance Dnn [defined as the position of the maxima of the radial distribution functions (D)] shows a nonmonotonic behavior as a function of the divalent ion concentration in the reservoir. Note that the shorter contour length of the DNA (~7.6 nm vs. ~22 nm in experiments) and smaller core (4.5 nm vs. 9 nm in experiments) in the simulations lead to notably smaller Dnn values. (F) The mean ionic concentrations in the dense phase as a function of the divalent ion concentration in the reservoir. Note that in the DNA-NP aggregate expansion regime (>2.0 M), the divalent ion concentration in the reservoir exceeds that in the DNA-NP dense phase. The dashed line signifies where the concentration in the dense phase is equal to that of the reservoir. These results were corroborated via chemical potential calculations (SI Appendix, Table S6).
The results of our MD simulations suggest that altered ionic distributions in the DNA shell and in the bulk solutions at high salinity underpin the observed lattice swelling. It is challenging to directly measure the ionic distributions surrounding DNA-NP via X-ray scattering because the scattered intensities are dominated by the electron-dense Au cores (41). However, WAXS can be used to measure Å to nm scale interionic structuring in the bulk solution. Thus, we prepared solutions of CaCl2, MgCl2, and SrCl2 of the same concentrations used in the assembly experiments and conducted in situ WAXS measurements. Background subtracted WAXS intensity profiles for CaCl2, MgCl2 and SrCl2 salt concentration series are shown in Fig. 4 A−C. The scattering peaks for q > 2 Å−1 correspond to water structure and ion hydration (42, 43). The present analysis will focus on q < 2 Å−1, which corresponds to ion ordering beyond local solvation.
Fig. 4.

Electrolyte ion–ion interactions studied with WAXS and varied solvent composition. (A−C) Capillary-subtracted WAXS intensity profiles for aqueous CaCl2 (A), SrCl2 (B), and MgCl2 (C) solutions. Two sets of peaks, which are not apparent in deionized water (blue profiles in A–C), are denoted as “Peak 1” and “Peak 2”. (D) Plot demonstrating the concentration-dependence of the peak positions, and their corresponding real-space separations. The position of Peak 1 scales with the cube root of concentration and is consistent with a repulsive interaction. Peak 2 does not shift appreciably with the salt concentrations but does shift to lower q (larger d spacing) with increasing cation size. Determination of peak positions is detailed in SI Appendix, Fig. S7. (E) Interparticle separation (Dnn) as a function of [CaCl2] in solutions with 0%, 5%, and 10% ethanol (v/v). Yellow markers denote precrystals, blue FCC, red BCC, and green RCP. Vertical lines mark the estimated salt concentrations corresponding to the minimum spacings observed in each sequence. The transition between contraction and expansion occurs more rapidly with higher ethanol volume fractions. Additionally, higher ethanol volume fractions lead to more contracted structures at low [CaCl2], and more expanded structures at high [CaCl2]. Corresponding SAXS patterns can be found in SI Appendix, Fig. S8. (F) A comparison of capillary-subtracted WAXS intensity profiles of CaCl2 solutions of various concentrations, with and without 10% ethanol. Notably, the peak at q ~ 1.5 Å (“Peak 2”) is enhanced in the presence of ethanol.
For all salt solutions, a low q intensity modulation (labeled “peak 1”) is observed. The corresponding intensity maxima shift monotonically to higher q, with a direct dependence on the cubic root of the salt concentration (, Fig. 4D). Based on previous studies on multivalent metal halide salts (44–46), peak 1 can be assigned to arise from repulsions between divalent cations. This is because the separation between the scattering ions decreases with increasing salt concentration () in a manner consistent with the distribution of repulsive objects in confinement in three dimensions. This is also verified by noting that depends on the salt concentration, but not on the cation type. In fact, the dependence for all salt solutions follows a “universal” curve (Fig. 4D). At higher salt concentrations (), where the DNA-NP structures expand with added salt, an additional modulation (labeled “peak 2”) near q ~1.5 Å−1 (d ~ 4 Å) becomes prominent (Fig. 4 A−C). The positions of this peak are independent of salt concentration but vary with cation type. In fact, the corresponding distance increases with cation size (Fig. 4D). It is concluded that peak 2 arises from attractive interactions between the divalent cations and anions. Previous scattering studies have also observed such intensity modulations (43, 46) and attributed them to positional correlations between hydrated cations and anions. This is further verified by the observations that the intensity of this peak is enhanced in salt solutions with anions of higher atomic number, e.g., CaBr2 vs. CaCl2 (45). For the specific case of CaCl2, additional evidence from dielectric relaxation spectroscopy (47) and Raman spectroscopy (48) support that Ca2+ and Cl− form complexes with shared hydration shells, called solvent-separated ion pairs.
The ionic ordering induced by attractive electrostatic interactions in our WAXS measurements provides a plausible explanation for the observed ion redistributions in the MD simulations. In particular, at high salt concentrations, cations are no longer independent. Complexation with anions increases the effective volumes occupied by the ions in the DNA-NP condensed phase and restricts the ability of the cations to organize densely in the DNA shells, thereby driving the local cation concentration below that of the bulk solution. Furthermore, the cation–anion pairing/correlations can effectively reduce the number of free cations that can interact with the DNA-NP, thus explaining the observed DNA-NP reexpansion at high salinities. That is, since increasing the number of counterions in the DNA-NP shell at low salt concentrations leads to compression of the assemblies, reducing the effective concentration of free counterions by clustering could lead to a reverse effect. Such a reduction in free ion concentration caused by the formation of electrolyte ion clusters has been recently proposed in theoretical treatments of concentrated electrolytes (49).
To further establish the role of ion–ion interaction in the aggregate reexpansion, we analyzed the DNA-NP assemblies in CaCl2 solutions with solvents of varied permittivity. Specifically, we utilized ethanol–water mixtures with 5% and 10% v/v ethanol. Ethanol and water are fully miscible and increasing ethanol content decreases the static dielectric constant of the mixture (50), thus increasing the electrostatic coupling within the solution (51). For example, enhanced interactions between DNA and ions in ethanol are known to lower the salt concentrations required to precipitate DNA (52). Therefore, in the low salt concentration regime, the enhanced DNA–ion attractions in ethanol should lower the minimum CaCl2 concentration required for DNA-NP crystallization. Furthermore, cation–anion interactions are enhanced in ethanol–water mixtures (53); based on our prior observations implicating ionic correlations as driving reexpansion, we hypothesized that increasing ethanol fractions will shift Cth to lower CaCl2 concentrations.
SAXS/WAXS measurements reveal the effects of enhanced DNA–ion and cation–anion attractive interactions in solutions with ethanol. With regard to DNA–ion interactions at low salt, we find that, as hypothesized, the CaCl2 concentration required for DNA-NP assembly into crystals decreases with increased ethanol content and that the boundaries for transitions to BCC and RCP structures shift to lower CaCl2 concentrations. For example, in a 5% ethanol solution, [CaCl2] ~ 0.175 M (vs. 0.3 M in 0% ethanol) is sufficient for forming DNA-NP FCC structures (Fig. 4E). Below Cth, for a given CaCl2 concentration, the Dnn is lowest for the 10% ethanol case and highest for the 0% ethanol case. By contrast, at high salinities ([CaCl2] > 3.5 M), i.e., in the regime where electrolyte cation–anion interactions drive lattice reexpansion, Dnn is found to increase with ethanol content (Fig. 4E). Furthermore, Cth shifts downward with increasing ethanol fraction: Cth ~ 3.3, 2.8, and 2.3 M for 0%, 5%, and 10% ethanol, respectively. This enhanced reexpansion and the above-described lowering of Cth are directly correlated to the enhanced propensity for short-ranged ordering of cations and anions in the solution as verified by WAXS (Fig. 4F). Specifically, WAXS intensity profiles exhibit higher amplitudes for the intensity modulations associated with Ca2+–Cl− ion paring (Peak 2) in 10% ethanol solutions, as compared to the 0% ethanol solutions, thus implying a higher degree of ionic clustering in the ethanol-containing solutions. These observations establish an empirical connection between the enhanced ion-pairing in bulk solution and the anomalous expansion of the DNA-NP aggregates in high salinities.
The effects of the ordered structure of electrolyte solution at high salinities have been invoked in other systems where nonmonotonic changes are observed. Most notably, in the aforementioned series of surface force measurements (10–12), an upturn in the range of electrostatic interactions (screening length) was observed (dubbed “underscreening”), which was attributed to strong positional correlations between the electrolyte cations and anions. It is also hypothesized that this underscreening effect is the cause of reentrant swelling and redissolution of polyelectrolytes at high salt concentrations (54). Parsimony suggests a unified explanation for all the nonmonotonic system responses: namely, underscreening, polyelectrolyte redissolution, gel reentrant swelling, and nanoparticle aggregate expansion observed at high salinities. However, some differences must be noted. The underscreening effect in 1:1 electrolytes (e.g., NaCl) was observed for . Here, a is the average diameters of the unhydrated electrolyte ions and is the classical electrostatic screening length. Extending this result to the present study, using the average of the unhydrated diameters of Ca2+ (0.20 nm) and Cl- (0.362 nm) (55), would imply Cth ~ 0.3 to 0.4 M for CaCl2, which is much smaller than the experimentally observed 3.3 M. Furthermore, the experimentally observed Cth sequence (Sr2+ < Mg2+ < Ca2+) seemingly depends on salt solubility limit, rather than ion size (SI Appendix, Fig. S5). Additional experiments on DNA-NP assembly, such as anomalous X-ray scattering measurements on low electron density cores (e.g., proteins) functionalized with DNA in specific salt solutions (e.g., SrBr2) could measure the composition and extent of the ionic atmosphere near these particles, thus, resolving ion–ion correlations at DNA-NP interfaces in the assembly contraction and expansion regimes.
To summarize, the present study demonstrates that ionic correlation at interfaces and in bulk solutions modulate interactions between charged macromolecules over the full range of achievable salinities, up to saturation. The assembly behavior of DNA-NP was analyzed concomitantly with the structure of the bulk electrolyte by combining X-ray scattering and MD simulations. Our study identifies two regimes. First, in solutions with salinities below half the solubility limit of the dissolved salts, divalent cations are free to organize around DNA molecules, inducing net attractive forces between DNA-NPs that drive the aggregation of DNA-nanoparticles. In this regime, adding salt strengthens DNA-NP interactions, causing assemblies to become more compact. Second, in the regime of salinities approaching the solubility limit, divalent cations form clusters with neighboring anions in the dense electrolyte solution. In this regime, these clusters partition favorably into the bulk solution away from the DNA-NP aggregates. Adding salt strengthens this electrolyte ordering, with the net effect of reducing the effective counterion concentration that can interact with the nanoparticles. This reverses the trend of aggregate compression with added salt. Strengthening electrolyte ion ordering by lowering solution permittivity shifts the transition between these two regimes to lower concentrations. Overall, our study provides insight on how ion–ion correlations shape interactions of electrolytes with charged materials at the nanoscale, and furthers the discussion on electrostatics in concentrated electrolytes.
Materials and Methods
Experimental Methods.
Thiol-terminated oligonucleotides (5′-T35-C3SH-3′) were synthesized using a MerMade solid-state controlled pore glass DNA synthesizer (BioAutomation) via phosphoramidite chemistry. Following separation from the pore glass, and purification via reverse-phase high-performance liquid chromatography (1260 Infinity II LC System, Agilent Technologies), the thiolated DNA was attached to the Au nanoparticles with a slow salting out procedure, with loading quantified via Quant-iT Oligreen Assay (Invitrogen). For X-ray scattering measurements, DNA-NP were dispersed in concentrated solutions of CaCl2 or MgCl2 or SrCl2 and allowed to equilibrate in sealed quartz capillaries for 1 to 2 d prior to measurements. X-ray scattering data were collected primarily at beamline 5ID-D of the Advanced Photon Source. X-ray energies of 16 and 17 keV were used. The SAXS/WAXS scattered intensities were measured simultaneously using three Rayonix CCD detectors positioned at 0.2 m, 1.0 m, and 7.5 m from the sample. Other details of sample preparation, X-ray measurements, data reduction, and analysis can be found in SI Appendix.
Simulation Methods.
We conducted MD simulations of a dense phase of nanoparticles functionalized with single-stranded DNA chains. The nanoparticles are embedded in an electrolyte solution which is in equilibrium with a reservoir. The chains, ions, and spherical nanoparticles are composed of spherical beads of the same size. The solvent is modeled as a uniform medium of dielectric constant εr. The beads interact via a truncated Lennard-Jones potential plus the Coulomb potential between charged beads. The system is equilibrated by performing 107 time-steps of MD simulation. The interparticle distances are calculated from a production run of at least 3 × 107 simulation time-steps. Further simulation details and the system parameters are reported in SI Appendix.
Supplementary Material
Appendix 01 (PDF)
Acknowledgments
This work was primarily supported by Department of Energy, Office of Science, Basic Energy Sciences award number DE-FG02-08ER46539. The SAXS/WAXS experiments were performed at the DuPont-Northwestern-Dow Collaborative Access Team (DND-CAT) located at Sector 5 of the Advanced Photon Source (APS) and at APS Sector 12. The APS, an Office of Science User Facility operated for DOE by Argonne National Laboratory, is supported by DOE under Contract DE-AC02-06CH11357. DND-CAT is supported by Northwestern University, The Dow Chemical Company, and DuPont de Nemours, Inc. This work made use of the Integrated Molecular Structure Education and Research Center facility at Northwestern University, which has received support from the Soft and Hybrid Nanotechnology Experimental Resource (NSF ECCS-2025633), the State of Illinois, and the International Institute for Nanotechnology. We thank Dr. Trung Dac Nguyen of UChicago, for helpful discussions, Soenke Seifert of Argonne National lab for assistance with SAXS measurements, and the Mirkin Group of Northwestern University for assistance in nucleic acid synthesis and characterization, in particular Jennifer Delgado, Kaitlin Landy, and Matthew Vasher.
Author contributions
R.J.E.R., S.K., F.J.-Á., M.J.B., and M.O.d.l.C. designed research; R.J.E.R., S.K., F.J.-Á., and S.J.W. performed research; R.J.E.R., S.K., F.J.-Á., M.J.B., and M.O.d.l.C. analyzed data; and R.J.E.R., S.K., F.J.-Á., M.J.B., and M.O.d.l.C. wrote the paper.
Competing interests
The authors F.J.-Á. and M.O.d.l.C. participated in a Faraday Discussion meeting on iontronics in June 2023, at Edinburgh, UK, which the reviewer S.P. and her group also attended. F.J.-Á. presented an article [Felipe Jiménez-Ángeles, Ali Ehlen, and Monica Olvera de la Cruz, Faraday Discuss. 246, 576–591 (2023)], and a member of the group of S.P. also presented an article in this meeting [Y. K. Catherine Fung and Susan Perkin, Faraday Discuss. 246, 370–386 (2023)]. The questions and answers on the presentations during all the different session in the meting are also published in Faraday Discuss. 2023, Vol. 246, under manuscripts with titles such as “Ionotronic coupling: general discussion” and “Iontronics under confinement: general discussion”, where F.J.-Á. participated on pages 592–617 and 157–178 and M.O.d.l.C. on pages 592–617, 466–486, and 322–355. These manuscripts are only reports on the discussions among the participants concerning articles presented during the meeting; For example, F.J.-Á. answered a short question by S.P. (in addition to 5 more questions by others) published on Vol. 246, pages 592–617, and M.O.d.l.C. asked a question to Christian Holm in the same session. Both F.J.-Á. and M.O.d.l.C. declare no conflict of interest with S.P.
Footnotes
Reviewers: V.S.C., Australian National University; and S.P., University of Oxford.
Data, Materials, and Software Availability
All data supporting this study are included in the article and/or SI Appendix.
Supporting Information
References
- 1.Zang X. N., Shen C. W., Sanghadasa M., Lin L. W., High-voltage supercapacitors based on aqueous electrolytes. ChemElectroChem 6, 976–988 (2019). [Google Scholar]
- 2.Wennerstrom H., Estrada E. V., Danielsson J., Oliveberg M., Colloidal stability of the living cell. Proc. Natl. Acad. Sci. U.S.A. 117, 10113–10121 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Debye P., Huckel E., The theory of electrolytes I. The lowering of the freezing point and related occurrences. Phys. Z. 24, 185–206 (1923). [Google Scholar]
- 4.Blum L., Hoye J. S., Mean spherical model for asymmetric electrolytes. 2. Thermodynamic properties and pair correlation-function. J. Phys. Chem. 81, 1311–1317 (1977). [Google Scholar]
- 5.McBride A., Kohonen M., Attard P., The screening length of charge-asymmetric electrolytes: A hypernetted chain calculation. J. Chem. Phys. 109, 2423–2428 (1998). [Google Scholar]
- 6.Gonzalez-Mozuelos P., Yeom M. S., Olvera de la Cruz M., Molecular multivalent electrolytes: Microstructure and screening lengths. Eur. Phys. J. E. 16, 167–178 (2005). [DOI] [PubMed] [Google Scholar]
- 7.Jing Y. F., Jadhao V., Zwanikken J. W., Olvera de la Cruz M., Ionic structure in liquids confined by dielectric interfaces. J. Chem. Phys. 143, 194508 (2015). [DOI] [PubMed] [Google Scholar]
- 8.Lee S. S., Koishi A., Bourg I. C., Fenter P., Ion correlations drive charge overscreening and heterogeneous nucleation at solid-aqueous electrolyte interfaces. Proc. Natl. Acad. Sc.i U.S.A. 118, e2105154118 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Zhang H., et al. , Stable colloids in molten inorganic salts. Nature 542, 328–331 (2017). [DOI] [PubMed] [Google Scholar]
- 10.Gebbie M. A., et al. , Ionic liquids behave as dilute electrolyte solutions. Proc. Natl. Acad. Sci. U.S.A. 110, 9674–9679 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Smith A. M., Lee A. A., Perkin S., The electrostatic screening length in concentrated electrolytes increases with concentration. J. Phys. Chem. Lett. 7, 2157–2163 (2016). [DOI] [PubMed] [Google Scholar]
- 12.Lee A. A., Perez-Martinez C. S., Smith A. M., Perkin S., Underscreening in concentrated electrolytes. Faraday Discuss. 199, 239–259 (2017). [DOI] [PubMed] [Google Scholar]
- 13.Li Y. H., Girard M., Shen M., Millan J. A., Olvera de la Cruz M., Strong attractions and repulsions mediated by monovalent salts. Proc. Natl. Acad. Sci. U.S.A. 114, 11838–11843 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Hjalmarsson N., Atkin R., Rutland M. W., Switchable long-range double layer force observed in a protic ionic liquid. Chem. Commun. 53, 647–650 (2017). [DOI] [PubMed] [Google Scholar]
- 15.Gaddam P., Ducker W., Electrostatic screening length in concentrated salt solutions. Langmuir 35, 5719–5727 (2019). [DOI] [PubMed] [Google Scholar]
- 16.Anousheh N., Solis F. J., Jadhao V., Ionic structure and decay length in highly concentrated confined electrolytes. AIP Adv. 10, 125312 (2020). [Google Scholar]
- 17.Zeman J., Kondrat S., Holm C., Ionic screening in bulk and under confinement. J. Chem. Phys. 155, 204501 (2021). [DOI] [PubMed] [Google Scholar]
- 18.Kumar S., et al. , Absence of anomalous underscreening in highly concentrated aqueous electrolytes confined between smooth silica surfaces. J. Colloid Interface Sci. 622, 819–827 (2022). [DOI] [PubMed] [Google Scholar]
- 19.Härtel A., Bültmann M., Coupette F., Anomalous underscreening in the restricted primitive model. Phys. Rev. Lett. 130, 108202 (2023). [DOI] [PubMed] [Google Scholar]
- 20.Kewalramani S., et al. , Electrolyte-mediated assembly of charged nanoparticles. ACS Cent. Sci. 2, 219–224 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Yang W. H., Schatz G. C., Vanduyne R. P., Discrete dipole approximation for calculating extinction and raman intensities for small particles with arbitrary shapes. J. Chem. Phys. 103, 869–875 (1995). [Google Scholar]
- 22.Finney J. L., Random packings and structure of simple liquids. 1. geometry of random close packing. Proc. R. Soc. Lond. A 319, 479–493 (1970). [Google Scholar]
- 23.Sirota E. B., et al. , Complete phase diagram of a charged colloidal system: A synchrotron x-ray scattering study. Phys. Rev. Lett. 62, 1524–1527 (1989). [DOI] [PubMed] [Google Scholar]
- 24.Tan S. J., et al. , Crystallization of DNA-capped gold nanoparticles in high-concentration, divalent salt environments. Angew. Chem. Int. Ed. 53, 1316–1319 (2014). [DOI] [PubMed] [Google Scholar]
- 25.Mcconnell G. A., Gast A. P., Huang J. S., Smith S. D., Disorder-order transitions in soft-sphere polymer micelles. Phys. Rev. Lett. 71, 2102–2105 (1993). [DOI] [PubMed] [Google Scholar]
- 26.Yun H., et al. , Symmetry transitions of polymer-grafted nanoparticles: Grafting density effect. Chem. Mater. 31, 5264–5273 (2019). [Google Scholar]
- 27.Reinertsen R. J., Jimenez-Angeles F., Kewalramani S., Bedzyk M. J., Olvera de la Cruz M., Transformations in crystals of DNA-functionalized nanoparticles by electrolytes. Faraday Discuss., 10.1039/D3FD00109A (2024). [DOI] [PubMed] [Google Scholar]
- 28.Samanta D., et al. , Multivalent cation-induced actuation of DNA-mediated colloidal superlattices. J. Am. Chem. Soc. 141, 19973–19977 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Srivastava S., Chhabra A., Gang O., Effect of mono- and multi-valent ionic environments on the in-lattice nanoparticle-grafted single-stranded DNA. Soft Matter 18, 526–534 (2022). [DOI] [PubMed] [Google Scholar]
- 30.Hansen J.-P., Verlet L., Phase transitions of the lennard-jones system. Phys. Rev. 184, 151–161 (1969). [Google Scholar]
- 31.Bai Y., et al. , Quantitative and comprehensive decomposition of the ion atmosphere around nucleic acids. J. Am. Chem. Soc. 129, 14981–14988 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Ahmad R., Arakawa H., Tajmir-Riahi H. A., A comparative study of DNA complexation with Mg(II) and Ca(II) in aqueous solution: Major and minor grooves bindings. Biophys. J. 84, 2460–2466 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Long M. P., Alland S., Martin M. E., Isborn C. M., Molecular dynamics simulations of alkaline earth metal ions binding to DNA reveal ion size and hydration effects. Phys. Chem. Chem. Phys. 22, 5584–5596 (2020). [DOI] [PubMed] [Google Scholar]
- 34.Olvera de la Cruz M., et al. , Precipitation of highly charged polyelectrolyte solutions in the presence of multivalent salts. J. Chem. Phys. 103, 5781–5791 (1995). [Google Scholar]
- 35.Raspaud E., Olvera de la Cruz M., Sikorav J. L., Livolant F., Precipitation of DNA by polyamines: A polyelectrolyte behavior. Biophys. J. 74, 381–393 (1998). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Nguyen T. T., Rouzina I., Shklovskii B. I., Reentrant condensation of DNA induced by multivalent counterions. J. Chem. Phys. 112, 2562–2568 (2000). [Google Scholar]
- 37.Besteman K., Van Eijk K., Lemay S., Charge inversion accompanies DNA condensation by multivalent ions. Nat. Phys. 3, 641–644 (2007). [Google Scholar]
- 38.Sing C. E., Zwanikken J. W., Olvera de la Cruz M., Effect of ion–ion correlations on polyelectrolyte gel collapse and reentrant swelling. Macromolecules 46, 5053–5065 (2013). [Google Scholar]
- 39.Solis F. J., Olvera de la Cruz M., Flexible linear polyelectrolytes in multivalent salt solutions: Solubility conditions. EPJ Direct 2, 1–18 (2000). [Google Scholar]
- 40.Rumble J., Ed., CRC Handbook of Chemistry and Physics (CRC Press, 2023), vol. 104. [Google Scholar]
- 41.Kewalramani S., et al. , Counterion distribution surrounding spherical nucleic acid-Au nanoparticle conjugates probed by small-angle X-ray scattering. ACS Nano 7, 11301–11309 (2013). [DOI] [PubMed] [Google Scholar]
- 42.Jalilehvand F., et al. , Hydration of the calcium ion. An EXAFS, large-angle X-ray scattering, and molecular dynamics simulation study. J. Am. Chem. Soc. 123, 431–441 (2001). [DOI] [PubMed] [Google Scholar]
- 43.Badyal Y. S., Barnes A. C., Cuello G. J., Simonson J. M., Understanding the effects of concentration on the solvation structure of Ca2+ in aqueous solutions. II: Insights into longer range order from neutron diffraction isotope substitution J. Phys. Chem. A 108, 11819–11827 (2004). [Google Scholar]
- 44.Marques M. A., Cabaco M. I., Marques M. I. D., Gaspar A. M., Intermediate-range order in aqueous solutions of salts constituted of divalent ions combined with monovalent counter-ions. J. Phys. Condens. Matter 14, 7427–7448 (2002). [Google Scholar]
- 45.Gaspar A. M., et al. , X-ray diffraction investigations of concentrated aqueous solutions of calcium halides. J. Mol. Liq. 110, 15–22 (2004). [Google Scholar]
- 46.Fetisov E. O., et al. , Nanometer-scale correlations in aqueous salt solutions. J. Phys. Chem. Lett. 11, 2598–2604 (2020). [DOI] [PubMed] [Google Scholar]
- 47.Friesen S., Hefter G., Buchner R., Cation hydration and ion pairing in aqueous solutions of MgCl2 and CaCl2. J. Phys. Chem. B 123, 891–900 (2019). [DOI] [PubMed] [Google Scholar]
- 48.Rudolph W. W., Irmer G., Hydration of the calcium(II) ion in an aqueous solution of common anions (ClO4-, Cl-, Br-, and NO3-). Dalton Trans. 42, 3919–3935 (2013). [DOI] [PubMed] [Google Scholar]
- 49.Safran S. A., Pincus P. A., Scaling perspectives of underscreening in concentrated electrolyte solutions. Soft Matter 19, 7907–7911 (2023). [DOI] [PubMed] [Google Scholar]
- 50.Wyman J., The dielectric constant of mixtures of ethyl alcohol and water from -5 to 40°. J. Am. Chem. Soc. 53, 3292–3301 (1931). [Google Scholar]
- 51.Laanait N., et al. , Tuning ion correlations at an electrified soft interface. Proc. Natl. Acad. Sci. U.S.A. 109, 20326–20331 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Arscott P. G., Ma C. L., Wenner J. R., Bloomfield V. A., DNA condensation by cobalt hexaammine(Iii) in alcohol-water mixtures—Dielectric-constant and other solvent effects. Biopolymers 36, 345–364 (1995). [DOI] [PubMed] [Google Scholar]
- 53.Bald A., Szejgis A., Barczyńska J., Piekarski H., Effect of ionic association on the B coefficient for CaCl2 in ethanol-water mixtures at 298.15 K. Phys. Chem. Liq. 37, 125–135 (1999). [Google Scholar]
- 54.Liu G. M., Parsons D., Craig V. S. J., Re-entrant swelling and redissolution of polyelectrolytes arises from an increased electrostatic decay length at high salt concentrations. J. Colloid Interface Sci. 579, 369–378 (2020). [DOI] [PubMed] [Google Scholar]
- 55.Shannon R. D., Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A. 32, 751–767 (1976). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Appendix 01 (PDF)
Data Availability Statement
All data supporting this study are included in the article and/or SI Appendix.

