Abstract
We explore the deconvolution of correlations for the interpretation of the microstructural behavior of aqueous electrolytes according to the neutron diffraction with isotopic substitution (NDIS) approach toward the experimental determination of ion coordination numbers of systems involving oxyanions, in particular, sulfate anions. We discuss the alluded interplay in the title of this presentation, emphasized the expectations, and highlight the significance of tackling the challenging NDIS experiments. Specifically, we focus on the potential occurrence of
Introduction
Oxyanions such as nitrates, carbonates, and sulfates are ubiquitous species in geochemical and industrial aqueous phase environments. In particular, the aqueous chemistry in the troposphere usually involves hydrated nitrates resulting from the fast interaction between nitric acid and aerosol particles [1], while their interactions with sulfates are crucial in the homogeneous nucleation of ice particles [2]. Moreover, sulfate ions trigger changes of electrode activity in fuel cells [3], and are usually the signature of water pollution in fresh water environments [4]. Yet, despite its environmental and industrial relevance, little is understood about the hydration structure of the sulfate ion. As in the case of the hydration of nitrates [5, 6], sulfates exhibit weak – though not to the extent of the monovalent oxyanions – interactions with water whose pair correlation functions typically overlap those of the cation–water interactions [7–10]. Its weak hydration behavior, added to the ionic strength effect on the water dielectric permittivity, provides the opportunity to participate in cation-oxyanion
The current state of affairs is likely the result of a common factor behind the hydration of oxyanions manifested as the overlapping of several inter- and intra-atomic correlation peaks within the radial distance, 2.0<r(Å)<4.5, including in this case the
Surprisingly, we are not aware of any NDIS data available in the literature for aqueous sulfates involving sulfur isotopic substitutions, despite the significant natS/33S neutron coherent scattering-length contrast [53]. The reason for this shortage might be traced back to the fact that the conventional combination of natS/33S and H/D substitutions provides complementary information for the solvation structure around sulfate anion, yet, neither addresses directly the near-neighbor coordination between the anion’s oxygen and water. In fact, the few NDIS available experiments involving metal sulfates of which we are aware actually targeted the metal hydration behavior rather than the anion [54–57].
In this context, the main goal of the current work is to illustrate how a judicious interplay between statistical mechanics theory and molecular simulation can be used to test the accuracy and internal consistency of the NDIS approach for the determination of the aqueous coordination of the sulfate anion, i.e., as a dry-run for the actual experiments. For that purpose, we explore the eventual deconvolution of the contributions from
Fundamentals
Neutron diffraction with isotopic substitution involving monoatomic ionic species has become a mature approach for the microstructural studies of aqueous electrolyte solutions [58–60] as long as the information from the NDIS raw data is corrected for the presence of ion-pairing [6, 61–64]. In contrast, NDIS experiments on aqueous electrolytes for the determination of the hydration microstructure of oxyanions involving dual αY/βY and natO/18O substitution are currently inexistent obviously due to the short time elapsed since Fischer et al.’s findings [65] regarding the more accurate assessment of the natO/18O difference of coherent scattering length, and the challenging nature of the NDIS experiments.
As we have recently highlighted it [5, 6], and due to the considerable experimental challenges behind the NDIS experiments involving null (i.e., neutron transparency to specific atoms) environment, it becomes extremely valuable to have available a molecular-based tool to conduct a priori analyses of the sought NDIS experiments. In this context, the interplay between statistical mechanics theory and molecular-based simulation affords a rigorous way to test unambiguously the accuracy of the approximations underlying the NDIS methodology. This is achievable because molecular simulation provides the full characterization of the system microstructure, i.e., not only the s(s+1)/2 pair correlation functions for a system involving s interacting sites (vide infra), but also the resulting NDIS output for any molecular model representing the system of interest. Precisely for this reason, the test of accuracy of the NDIS formalism becomes independent of the choice of the interaction potential models used in the simulation, yet, the reader might be understandably interested in knowing how realistic the used models are for the description of the aqueous electrolytes under investigation, an issue we have already addressed in great detail elsewhere [6].
The access to the s(s+1)/2 pair correlation functions by molecular simulation provides the unmatched opportunity to locate the relevant diffraction peaks, the presence of peak overlapping, and ultimately to facilitate the interpretation of the diffraction data. This is especially relevant for the study of oxyanion hydration, where it becomes practically impossible to extract any coordination information associated with the oxygen site of the oxyanion, unless we are able to discriminate it from the corresponding water–oxygen correlations. Therefore, in what follows we illustrate what we could expect to observe in a simulated NDIS experiment involving natO/18O and natS/33S in aqueous sulfate solutions, and discuss the implications on the accurate determination of water–sulfate coordination, where all microstructural details are known simultaneously with the corresponding first-order differences of the neutron weighted distribution functions. This feature allows the unambiguous identification of peak overlapping between pair correlation functions, and consequently, the isolation of specific pair correlation peaks for the assessment of meaningful species coordination as well as the opportunity to test conjectured behaviors underlying the potential occurrence of ion-pairing and its effect in the interpretation of the NDIS raw data [62, 66].
Neutron diffraction with dual natS/33S and natOS/18OS isotopic substitution
Here we examine the link between the measured quantities and the targeted microstructures (e.g., ion coordination environment), and identify the (frequently dismissed) problems underlying the meaningful interpretation of the experimental evidence, such as the unavoidable peak overlapping as the signature of ion-pair formation [62, 66]. The starting point is the portion of the neutron scattering differential cross section of the aqueous sample, dσ/dΩ, that comprises the desired information on the microstructure, i.e., the total structure factor F(k) defined as [67],
where k=(4π/λ)sin(θ/2), λ of the incident neutron wavelength, while θ, ci and bi denote the scattering angle, the atomic fraction and the coherent neutron scattering length of the atomic species i, respectively. Obviously F(k) is a linear combination of the partial structure factor Sij(k) describing the correlation between atoms of types i and j, i.e.,
where ρ and gij(r) are the atomic number density of the solution and the corresponding radial pair distribution function for ij-pair interactions. For most practical purposes we prefer dealing with the total (real space neutron-weighted) pair correlation function G(r) rather than F(k), i.e., its Fourier transform,
Considering that a simple
The reasonable way to circumvent this shortcoming is through the cancelation of the undesired correlations via the implementation of the first-order difference scheme [69], i.e., comprising two diffraction experiments involving a pair of identical solutions except for the isotopic states of the species (labeled α or β) under study. Then, the ion coordination can be determined by the integration of the resulting first-order difference within the radial interval [rL, rU ] where the sought ion-water correlation occurs and exhibits no peak overlapping (vide infra). Specifically, the cancelation of contributions from the solvent–solvent and solvent–(non-substituted) solute species interactions in the first-order difference scheme for the study of the coordination around the
with
with
A further manipulation of the system environment could be achieved by making the M-cation transparent to the neutrons by means of the so-called null M-cation (nM) environment, involving the αM/βM isotopic substitution at an isotopic composition cαM/cβM=–(bβM/bαM ) that makes DS = DO = 0, so that eqs. (4)–(5) reduce to the following working expressions,
with
with

Atomistic peek of a representative contact

Normalized first-order difference of neutron weighted distribution functions for the 1.72 m NiSO4 in heavy- and null-water solution at ambient conditions under natS/33S substitution.

Normalized first-order difference of neutron weighted distribution functions for the 1.72 m NiSO4 in heavy- and null-water solution at ambient conditions under natO/18O substitution.
Typically, assuming that there are additional contributions (other than gDS(r)) to
where β=H, O and rs typically locates the first valley of the radial distribution function gβI(r), we have that the coordination of water around the sulfate’s sulfur site becomes,
and,
while the result corresponding to the sulfate’s oxygen site reads,
and,
Interaction potential models and molecular simulation methodology
For the analysis of the hydration behavior of the aqueous NiSO4 and to illustrate the interplay as well as issues discussed above, we performed isochoric–isothermal molecular dynamics simulations at ambient conditions. For that purpose we have chosen simple but reliable intermolecular potential models including the rigid SPC/E water [70], and the Wallen et al. [71], and the five-sites rigid tetrahedral Cannon et al. [72] for the nickel and sulfate ions, respectively. All these potential models involve Lennard–Jones models, with the unlike pair interactions described by the Lorentz–Berthelot combining rules, and site-site electrostatic interactions according to the force field collected in Table 1.
Potential parameters for the aqueous NiSO4 model solutions.c
| ii-Interaction | ε ii /k(K)a | σii(Å)b | q i (e) | Refs. |
|---|---|---|---|---|
| Ni···Ni | 50.34 | 2.050 | 2.0 | [71] |
| OS···OS | 30.09 | 3.150 | –1.100 | [72] |
| S···S | 125.7 | 3.550 | 2.400 | [72] |
| OW···OW | 125.7 | 3.166 | –0.8476 | [70] |
| H···H | – | – | 0.4238 | [70] |
a
All simulations involved NW = 1986 water molecules and Nions=2NNi=124 ions and were carried out according to our own implementation of a Nosé–Poincare algorithm [73, 74] for the integration of the Newton–Euler equations of motion within a cubic simulation box of dimension L under 3D periodic boundary conditions. The initial configuration was generated by the Packmol utility [75] and equilibrated by at least 1.0 ns, followed by production runs of 4.0 ns, using an integration time-step of 2.0 fs, during which the microstructural information was collected. All Lennard–Jones interactions were truncated at a cut-off radius min(rc ≈3.5σSPCE, rc=L/2), while the electrostatic interactions were handled by an Ewald summation whose convergence parameters were chosen to assure an error smaller than 5×10–5εSPCE for both the real and reciprocal spaces [76].
Microstructural results
From the ten simulated pair correlation functions we determined the first-order differences of neutron-weighted radial distribution functions, for the heavy-water –
In Table 2 we collect the values of the coherent scattering lengths used in the study, which were taken from the available literature. The first-order differences
Coherent neutron scattering lengths for aqueous NiSO4.
| Species | b coh(fm) | Refs. |
|---|---|---|
| natS | 2.847 | [53] |
| 33S | 4.74 | [53] |
| 1H | –3.74 | [53] |
| 2H | 6.67 | [53] |
| natO | 5.805 | [53] |
| 18O | 6.009 | [65] |
| natNi | 10.30 | [53] |
| 62Ni | –8.7 | [53] |
Coefficients of the first-order difference of weighted distributions for the natS/33S substitution in 1.72 m NiSO4 heavy- and null-aqueous solutions.
| Coefficienta | Heavy-water | Null-water |
|---|---|---|
| A S | 6.945079E-02 | 6.945079E-02 |
| B S | 0.1595992 | 5.732689E-02 |
| C S | 0.0 | –5.732689E-02 |
| D S | 3.970990E-03 | 3.970990E-03 |
| E S | 1.462520E-03 | 1.462520E-03 |
| F S | 8.952077E-03 | 8.952077E-03 |
|
|
0.2434356 | 8.383637E-02 |
aIn fm2 units.
Coefficients of the first-order difference of weighted distributions for the natOS/18OS substitution in 1.72 m NiSO4 heavy- and null-aqueous solutions.
| Coefficienta | Heavy-water | Null-water |
|---|---|---|
| A O | 2.993758E-02 | 2.993758E-02 |
| B O | 6.879714E-02 | 2.472620E-02 |
| C O | 0.0 | –2.472620E-02 |
| D O | 1.711742E-03 | 1.711742E-03 |
| E O | 3.926700E-03 | 3.926700E-03 |
| F O | 4.731388E-04 | 4.731388E-04 |
|
|
0.1048463 | 3.604916E-02 |
aIn fm2 units.
The main features of
However, the first peak of
First water coordination of the sulfate ion in 1.72 m NiSO4 heavy- and null-aqueous solutions according to the direct and the NDIS-based expressions.
| α–β interactions | Direct integral (ru)c | NDIS integral (ru)d |
|---|---|---|
| OW···S | 14.1(4.51 Å) | 14.7(4.43 Å)b |
| D···S | 11.6(3.40 Å) | 11.4(3.32 Å)a |
| OW···OS | 3.2(3.16 Å) | 3.6(3.24 Å)b |
| D···OS | 2.8(2.37 Å) | 3.1(2.37 Å)a |
| Ni···S | 0.4(3.37 Å) | – |
| OS···Ni | 0.4(2.20 Å) | 0.4(2.20 Å)b |
aHeavy-water; bnull-water; c
The first peak of
will experience the same problem as that for
Note that
In Fig. 3a–b we display the corresponding first-order differences
Then the actual
from eq. (12) to attain,
Obviously, according to Figs. 2 and 3, there are no chances of assessing the
so that,
However, we still have another source of information to characterize further the water coordination of the sulfate anion, i.e., the null-nickel venue through the manipulation of the isotopic composition of the cation. Under this null-nickel environment the relevant distribution functions

Normalized first-order difference of neutron weighted distribution functions for the 1.72 m NiSO4 in heavy-water null-Ni solution at ambient conditions under either natO/18O or natS/33S substitution.
and,
Note that due to the absence of any contact (anion site-cation) pair corrections, the coordination number determined by eq. (19) will provide a test of consistency for the calculations involving eq. (16). As an illustration of the issue, in Table 6 we made the comparison between the reference coordination numbers given by their statistical mechanical definition, i.e., eq. (8).
First water coordination of the sulfate ion in 1.72 m NiSO4 aqueous solutions from integrals over
| α–β Interactions | Direct integral(ru)b | NDIS integral (ru)a | NDIS integral (ru) |
|---|---|---|---|
| D···S | 11.6(3.40 Å) | 11.4(3.32 Å) | 11.4(3.32 Å)c |
| D···OS | 2.8(2.37 Å) | 3.1(2.37 Å) | 2.8(3.7 Å)d |
Up to this point we have presented the tools and illustrated their use not only to detect the presence of (CIP)
The degree of either
where P–(r) (P+(r)) denotes the probability that either
under the condition that, in the thermodynamic limit,
In Fig. 5a–b we display the
![Fig. 5:
Radial pair distribution function gNiS(r), the corresponding Pourier-deLap
N
i
+
2
⋯
S[O
4
2
−
]
$N{i^{ + 2}} \cdots {\rm{S[O}}_4^{2 - }]$
pair distribution function and the resulting degree of pair association for the 1.72 m NiSO4 aqueous solution at ambient conditions.](/document/doi/10.1515/pac-2015-1002/asset/graphic/j_pac-2015-1002_fig_032.jpg)
Radial pair distribution function gNiS(r), the corresponding Pourier-deLap
Discussion and final remarks
We have described the fundamentals underlying the interpretation of microstructural behavior of aqueous electrolytes according to the NDIS approach toward the experimental determination of ion coordination numbers of systems involving sulfate anions. We discussed the ‘philosophy’ motivating the interplay alluded to in the title of this presentation, and emphasized the expectations as well as significance of tackling the challenging NDIS experiments. Specifically, we highlighted the potential occurrence of
We must emphasize that this interplay becomes a formidable asset after recognizing the inherent molecular simulation ability to provide all pair correlation functions that fully characterize the system microstructure and allows us to “reconstruct” the eventual NDIS output, i.e., to take an atomistic “peek” at the local environment around the isotopically-labeled species before any experiment is ever attempted, and ultimately, to test the accuracy of the “measured” NDIS-based coordination numbers against the actual values by the “direct” counting. In addition, the isotopic differentiation between OW and OS in eqs. (4)–(7) provides a handy way to check the consistency of the experimental raw data sets according to the natural constraint given by the intra-molecular coordination
where
with
where gSOS(r)=gOSS(r). Note that the equality of the two lines in eqs. (23)–(25) is guaranteed as long as there is no overlapping of peaks within the range of intra-molecular interactions, i.e., the self-consistency of
where the identity (26) becomes an indicator of properly normalized experimental first-order differences
In summary, in developing the current interplay we have exposed some frequently overlooked issues and highlighted that the pressing challenge at this juncture is to confront the significant difficulties associated with these null-environment NDIS experiments and translate the proposed novel schemes into versatile tools for the accurate full characterization of oxyanion hydration.
This manuscript has been authored by UT-Battelle, LLC under Contract No. DE-AC05-00OR22725 with the U.S. Department of Energy. The United States Government retains and the publisher, by accepting the article for publication, acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or reproduce the published form of this manuscript, or allow others to do so, for United States Government purposes. The Department of Energy will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan).
Article note
A collection of invited papers based on presentations at the 34th International Conference on Solution Chemistry (ICSC-34), Prague, Czech Republic, 30th August – 3rd September 2015.
Acknowledgments
This work was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Chemical Sciences, Geosciences, and Biosciences Division.
References
[1] B. J. Finlayson-Pitts. Chem. Rev.103, 4801 (2003).10.1021/cr020653tSuche in Google Scholar PubMed
[2] V. Ramanathan, P. J. Crutzen, J. T. Kiehl, D. Rosenfeld. Science294, 2119 (2001).10.1126/science.1064034Suche in Google Scholar PubMed
[3] N. Hoshi, M. Kuroda, T. Ogawa, O. Koga, Y. Hori. Langmuir20, 5066 (2004).10.1021/la036149gSuche in Google Scholar PubMed
[4] G. A. Ulrich, G. N. Breit, I. M. Cozzarelli, J. M. Suflita. Env. Sci. Tec.37, 1093 (2003).10.1021/es011288aSuche in Google Scholar PubMed
[5] A. A. Chialvo, L. Vlcek. J. Phys. Chem. B119, 519 (2015).10.1021/acs.jpcb.5b00595Suche in Google Scholar PubMed
[6] A. A. Chialvo, L. Vlcek. Fluid Phase Equilib.407, 84 (2016).10.1016/j.fluid.2015.05.014Suche in Google Scholar
[7] R. Caminiti. J. Chem. Phys.84, 3336 (1986).10.1063/1.450268Suche in Google Scholar
[8] R. Caminiti. Chem. Phys. Lett.88, 103 (1982).10.3406/roma.1982.2101Suche in Google Scholar
[9] G. Licheri, G. Paschina, G. Piccaluga, G. Pinna. J. Chem. Phys.81, 6059 (1984).10.1063/1.447609Suche in Google Scholar
[10] M. Magini, G. Licheri, G. Paschina, G. Piccaluga, G. Pinna. X-Ray Diffraction of Ions in Aqueous Solutions: Hydration and Complex Formation, CRC Press, Inc., Boca Raton (1988).Suche in Google Scholar
[11] L. G. Sillén, I. U. o. Pure, A. C. C. o. E. Data, A. E. Martell. Stability Constants of Metal-ion Complexes [with] Supplement. Chemical Society (1971).Suche in Google Scholar
[12] S. Katayama. Bull. Chem. Soc. Jpn.46, 106 (1973).Suche in Google Scholar
[13] F. H. Fisher, A. P. Fox. J. Solution Chem.8, 309 (1979).10.1007/BF00650748Suche in Google Scholar
[14] H. Yokoyama, T. Ohta. Bull. Chem. Soc. Jpn.62, 345 (1989).10.1246/bcsj.62.345Suche in Google Scholar
[15] M. Tomsic, M. Bester-Rogac, A. Jamnik, R. Neueder, J. Barthel. J. Solution Chem.31, 19 (2002).10.1023/A:1014853001357Suche in Google Scholar
[16] M. Bester-Rogac, V. Babic, T. M. Perger, R. Neueder, J. Barthel. J. Mol. Liq.118, 111 (2005).10.1016/j.molliq.2004.07.023Suche in Google Scholar
[17] M. Bester-Rogac. J. Chem. Eng. Data53, 1355 (2008).10.1021/je8001255Suche in Google Scholar
[18] M. Madekufamba, P. R. Tremaine. J. Chem. Eng. Data56, 889 (2011).10.1021/je100729tSuche in Google Scholar
[19] R. Buchner, S. G. Capewell, G. Hefter, P. M. May. J. Phys. Chem. B103, 1185 (1999).10.1021/jp983706cSuche in Google Scholar
[20] W. Wachter, S. Fernandez, R. Buchner, G. Hefter. J. Phys. Chem. B111, 9010 (2007).10.1021/jp072425eSuche in Google Scholar PubMed
[21] R. Buchner, T. Chen, G. Hefter. J. Phys. Chem. B108, 2365 (2004).10.1021/jp034870pSuche in Google Scholar
[22] T. Chen, G. Hefter, R. Buchner. J. Solution Chem.34, 1045 (2005).10.1007/s10953-005-6993-5Suche in Google Scholar
[23] G. Hefter. Pure Appl. Chem.78, 1571 (2006).10.1351/pac200678081571Suche in Google Scholar
[24] D. B. Bechtold, G. Liu, H. W. Dodgen, J. P. Hunt. J. Phys. Chem.82, 333 (1978).10.1021/j100492a019Suche in Google Scholar
[25] F. P. Daly, D. R. Kester, C. W. Brown. J. Phys. Chem.76, 3664 (1972).10.1021/j100668a027Suche in Google Scholar
[26] A. R. Davis, B. G. Oliver. J. Phys. Chem.77, 1315 (1973).10.1021/j100629a028Suche in Google Scholar
[27] F. Rull, C. Balarew, J. L. Alvarez, F. Sobron, A. Rodriguez. J. Raman Spectrosc.25, 933 (1994).10.1002/jrs.1250251206Suche in Google Scholar
[28] J. D. Frantz, J. Dubessy, B. O. Mysen. Chem. Geol.116, 181 (1994).10.1016/0009-2541(94)90013-2Suche in Google Scholar
[29] W. Rudolph, G. Irmer. J. Solution Chem.23, 663 (1994).10.1007/BF00972713Suche in Google Scholar
[30] W. W. Rudolph, G. Irmer, G. T. Hefter. PCCP5, 5253 (2003).10.1039/b308951gSuche in Google Scholar
[31] D. Watanabe, H. Hamaguchi. J. Chem. Phys.123, Article ID 34508 (2005).10.1063/1.1931660Suche in Google Scholar PubMed
[32] E. V. Petrova, M. A. Vorontsova, V. L. Manomenova, L. N. Rashkovich. Crystallogr. Rep.57, 579 (2012).10.1134/S1063774512010099Suche in Google Scholar
[33] W. D. Bale, E. W. Davies, C. B. Monk. T. Faraday Soc.52, 816 (1956).10.1039/tf9565200816Suche in Google Scholar
[34] R. Nasanen. Acta Chem. Scand.3, 179 (1949).10.3891/acta.chem.scand.03-0179Suche in Google Scholar
[35] H. Yokoyama, H. Yamatera. Bull. Chem. Soc. Jpn.48, 2719 (1975).10.1246/bcsj.48.2719Suche in Google Scholar
[36] P. A. Bergstrom, J. Lindgren, O. Kristiansson. J. Phys. Chem.95, 8575 (1991).10.1021/j100175a031Suche in Google Scholar
[37] J. Stangret, T. Gampe. J. Phys. Chem. A106, 5393 (2002).10.1021/jp014063vSuche in Google Scholar
[38] V. S. K. Nair, G. H. Nancollas. J. Chem. Soc. Article ID 791, 3934 (1959).10.1039/jr9590003934Suche in Google Scholar
[39] S. Kratsis, G. Hefter, P. M. May. J. Solution Chem.30, 19 (2001).10.1023/A:1005201825524Suche in Google Scholar
[40] C. Akilan, P. M. May, G. Hefter. J. Solution Chem.43, 885 (2014).10.1007/s10953-014-0170-7Suche in Google Scholar
[41] H. Yokoyama, H. Yamatera. Bull. Chem. Soc. Jpn.48, 2708 (1975).10.1246/bcsj.48.1770Suche in Google Scholar
[42] W. Libus, T. Sadowska, Z. Libus. J. Solution Chem.9, 341 (1980).10.1007/BF00651541Suche in Google Scholar
[43] R. Caminiti. Z. Naturforsch. A36, 1062 (1981).10.1515/zna-1981-1007Suche in Google Scholar
[44] R. Caminiti, G. Johansson. Acta Chem. Scand. A35, 373 (1981).10.3891/acta.chem.scand.35a-0373Suche in Google Scholar
[45] R. Caminiti, G. Paschina. Chem. Phys. Lett.82, 487 (1981).10.1016/0009-2614(81)85425-5Suche in Google Scholar
[46] G. Licheri, G. Paschina, G. Piccaluga, G. Pinna. Z. Naturforsch. A37, 1205 (1982).10.1515/zna-1982-1012Suche in Google Scholar
[47] T. Yamaguchi, O. Lindqvist. Acta Chem. Scand. A36, 377 (1982).10.3891/acta.chem.scand.36a-0377Suche in Google Scholar
[48] A. Musinu, G. Paschina, G. Piccaluga, M. Magini. J. Appl. Crystallogr.15, 621 (1982).10.1107/S0021889882012795Suche in Google Scholar
[49] T. Radnai, G. Palinkas, R. Caminiti. Z. Naturforsch. A37, 1247 (1982).10.1515/zna-1982-1105Suche in Google Scholar
[50] H. Ohtaki, T. Yamaguchi, M. Maeda. Bull. Chem. Soc. Jpn.49, 701 (1976).10.1246/bcsj.49.701Suche in Google Scholar
[51] V. Vchirawongkwin, B. M. Rode, I. Persson. J. Phys. Chem. B111, 4150 (2007).10.1021/jp0702402Suche in Google Scholar PubMed
[52] J. X. Xu, Y. Fang, C. H. Fang. Chin. Sci. Bull.54, 2022 (2009).10.1007/s11434-009-0232-1Suche in Google Scholar
[53] V. F. Sears. Neutron News3, 29 (1992).10.1080/10448639208218770Suche in Google Scholar
[54] S. Cummings. J. Phys. Paris45, 131 (1984).10.1051/jphyscol:1984713Suche in Google Scholar
[55] G. W. Neilson, J. E. Enderby. J. Phys. Chem.100, 1317 (1996).10.1021/jp951490ySuche in Google Scholar
[56] I. Howell, G. W. Neilson. J. Mol. Liq.73-4, 337 (1997).10.1016/S0167-7322(97)00077-9Suche in Google Scholar
[57] P. E. Mason, S. Ansell, G. W. Neilson, S. B. Rempe. J. Phys. Chem. B119, 2003 (2015).10.1021/jp511508nSuche in Google Scholar PubMed
[58] J. E. Enderby. Chem. Soc. Rev.24, 159 (1995).10.1039/cs9952400159Suche in Google Scholar
[59] G. W. Neilson, P. E. Mason, S. Ramos, D. Sullivan. Philos. T. Roy. Soc. A359, 1575 (2001).10.1098/rsta.2001.0866Suche in Google Scholar
[60] S. Ansell, A. C. Barnes, P. E. Mason, G. W. Neilson, S. Ramos. Biophys. Chem.124, 171 (2006).10.1016/j.bpc.2006.04.018Suche in Google Scholar PubMed
[61] Y. S. Badyal, A. C. Barnes, G. J. Cuello, J. M. Simonson. J. Phys. Chem. A108, 11819 (2004).10.1021/jp046476cSuche in Google Scholar
[62] A. A. Chialvo, J. M. Simonson. J. Chem. Phys.119, 8052 (2003).10.1063/1.1610443Suche in Google Scholar
[63] J. L. Fulton, S. M. Heald, Y. S. Badyal, J. M. Simonson. J. Phys. Chem. A107, 4688 (2003).10.1021/jp0272264Suche in Google Scholar
[64] T. Megyes, I. Bako, S. Balint, T. Grosz, T. Radnai. J. Mol. Liq.129, 63 (2006).10.1016/j.molliq.2006.08.013Suche in Google Scholar
[65] H. E. Fischer, J. M. Simonson, J. C. Neuefeind, H. Lemmel, H. Rauch, A. Zeidler, P. S. Salmon. J. Phys.-Condens. Mat.24, Article ID 505105 (2012).10.1088/0953-8984/24/50/505105Suche in Google Scholar PubMed
[66] A. A. Chialvo, J. M. Simonson. J. Chem. Phys.124, 154509 (2006).10.1063/1.2186641Suche in Google Scholar PubMed
[67] J. E. Enderby, G. W. Neilson. Rep. Prog. Phys.44, 593 (1981).10.1088/0034-4885/44/6/001Suche in Google Scholar
[68] D. L. Price, A. Pasquarello. Phys. Rev. B59, 5 (1999).10.1103/PhysRevB.59.5Suche in Google Scholar
[69] A. K. Soper, G. W. Neilson, J. E. Enderby. J. Phys. C. Solid State10, 1793 (1977).10.1088/0022-3719/10/11/014Suche in Google Scholar
[70] H. J. C. Berendsen, J. R. Grigera, T. P. Straatsma. J. Phys. Chem.91, 6269 (1987).10.1021/j100308a038Suche in Google Scholar
[71] S. L. Wallen, B. J. Palmer, J. L. Fulton. J. Chem. Phys.108, 4039 (1998).10.1063/1.475838Suche in Google Scholar
[72] W. R. Cannon, B. M. Pettitt, J. A. McCammon. J. Phys. Chem.98, 6225 (1994).10.1021/j100075a027Suche in Google Scholar
[73] S. Nose. J. Phys. Soc. Jpn.70, 75 (2001).10.1143/JPSJ.70.75Suche in Google Scholar
[74] H. Okumura, S. G. Itoh, Y. Okamoto. J. Chem. Phys.126, 084103 (2007).10.1063/1.2434972Suche in Google Scholar PubMed
[75] J. M. Martinez, L. Martinez. J. Comput. Chem.24, 819 (2003).10.1002/jcc.10216Suche in Google Scholar PubMed
[76] D. Fincham. Mol. Simulat.13, 1 (1994).10.1080/08927029408022180Suche in Google Scholar
[77] A. A. Chialvo, P. T. Cummings. J. Phys. Chem.100, 1309 (1996).10.1021/jp951445qSuche in Google Scholar
[78] M. Eigen, K. Tamm. Z. Elektrochem.66, 93 (1962).10.2307/310737Suche in Google Scholar
[79] J. C. Poirier, J. H. DeLap. J. Chem. Phys.35, 213 (1961).10.1063/1.1731893Suche in Google Scholar
[80] A. A. Chialvo, J. M. Simonson. Collec. Czech. Chem. Commun.75, 405 (2010).10.1135/cccc2009535Suche in Google Scholar
[81] S. Wasylkiewicz. Fluid Phase Equilib.57, 277 (1990).10.1016/0378-3812(90)85127-VSuche in Google Scholar
©2016 IUPAC & De Gruyter. This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License. For more information, please visit: http://creativecommons.org/licenses/by-nc-nd/4.0/
Artikel in diesem Heft
- Frontmatter
- Editorial
- Winners of the 2015 IUPAC-SOLVAY International Award for Young Chemists
- Conference papers
- “Thought experiments” as dry-runs for “tough experiments”: novel approaches to the hydration behavior of oxyanions
- Shedding light on the hydrophobicity puzzle
- From solutions to molecular emulsions
- New opportunities in the stereoselective dearomatization of indoles
- Chemoselective access to substituted butenolides via a radical cyclization pathway: mechanistic study, limits and application
- 1,3-Oxazinan-2-ones via carbonate chemistry: a facile, high yielding synthetic approach
- IUPAC Recommendations
- Glossary of terms used in computational drug design, part II (IUPAC Recommendations 2015)
- IUPAC Technical Reports
- Atomic weights of the elements 2013 (IUPAC Technical Report)
- Isotopic compositions of the elements 2013 (IUPAC Technical Report)
Artikel in diesem Heft
- Frontmatter
- Editorial
- Winners of the 2015 IUPAC-SOLVAY International Award for Young Chemists
- Conference papers
- “Thought experiments” as dry-runs for “tough experiments”: novel approaches to the hydration behavior of oxyanions
- Shedding light on the hydrophobicity puzzle
- From solutions to molecular emulsions
- New opportunities in the stereoselective dearomatization of indoles
- Chemoselective access to substituted butenolides via a radical cyclization pathway: mechanistic study, limits and application
- 1,3-Oxazinan-2-ones via carbonate chemistry: a facile, high yielding synthetic approach
- IUPAC Recommendations
- Glossary of terms used in computational drug design, part II (IUPAC Recommendations 2015)
- IUPAC Technical Reports
- Atomic weights of the elements 2013 (IUPAC Technical Report)
- Isotopic compositions of the elements 2013 (IUPAC Technical Report)