ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
Recently Viewed
You have not visited any articles yet, Please visit some articles to see contents here.
CONTENT TYPES
RETURN TO ISSUEPREVResearch ArticleNEXT

A Telescoping View of Solute Architectures in a Complex Fluid System

  • Ryuhei Motokawa*
    Ryuhei Motokawa
    Materials Sciences Research Center, Japan Atomic Energy Agency, Tokai, Ibaraki 319-1195, Japan
    *E-mail: [email protected]
  • Tohru Kobayashi
    Tohru Kobayashi
    Materials Sciences Research Center, Japan Atomic Energy Agency, Tokai, Ibaraki 319-1195, Japan
  • Hitoshi Endo
    Hitoshi Endo
    Materials Sciences Research Center, Japan Atomic Energy Agency, Tokai, Ibaraki 319-1195, Japan
    Neutron Science Division, Institute of Materials Structure Science, and Materials and Life Science Division, J-PARC Center, High Energy Accelerator Research Organization, 203-1 Shirakata, Tokai, Ibaraki 319-1106, Japan
    Department of Materials Structure Science, The Graduate University for Advanced Studies (SOKENDAI), 203-1 Shirakata, Tokai, Ibaraki 319-1106, Japan
    More by Hitoshi Endo
  • Junju Mu
    Junju Mu
    School of Chemical Engineering and Analytical Science, The University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
    More by Junju Mu
  • Christopher D. Williams
    Christopher D. Williams
    School of Chemical Engineering and Analytical Science, The University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
  • Andrew J. Masters*
    Andrew J. Masters
    School of Chemical Engineering and Analytical Science, The University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
    *E-mail: [email protected]
  • Mark R. Antonio
    Mark R. Antonio
    Chemical Sciences & Engineering Division, Argonne National Laboratory, Lemont, Illinois 60439, United States
  • William T. Heller
    William T. Heller
    Neutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United States
  • , and 
  • Michihiro Nagao
    Michihiro Nagao
    NIST Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg, Maryland 20899-6102, United States
    Center for Exploration of Energy and Matter, Department of Physics, Indiana University, Bloomington, Indiana 47408, United States
Cite this: ACS Cent. Sci. 2019, 5, 1, 85–96
Publication Date (Web):December 31, 2018
https://doi.org/10.1021/acscentsci.8b00669
Copyright © 2018 American Chemical Society
ACS AuthorChoiceACS AuthorChoicewith CC-BYlicense
Article Views
3028
Altmetric
-
Citations
LEARN ABOUT THESE METRICS
PDF (3 MB)
Supporting Info (1)»

Abstract

Short- and long-range correlations between solutes in solvents can influence the macroscopic chemistry and physical properties of solutions in ways that are not fully understood. The class of liquids known as complex (structured) fluids—containing multiscale aggregates resulting from weak self-assembly—are especially important in energy-relevant systems employed for a variety of chemical- and biological-based purification, separation, and catalytic processes. In these, solute (mass) transfer across liquid–liquid (water, oil) phase boundaries is the core function. Oftentimes the operational success of phase transfer chemistry is dependent upon the bulk fluid structures for which a common functional motif and an archetype aggregate is the micelle. In particular, there is an emerging consensus that mass transfer and bulk organic phase behaviors—notably the critical phenomenon of phase splitting—are impacted by the effects of micellar-like aggregates in water-in-oil microemulsions. In this study, we elucidate the microscopic structures and mesoscopic architectures of metal-, water-, and acid-loaded organic phases using a combination of X-ray and neutron experimentation as well as density functional theory and molecular dynamics simulations. The key conclusion is that the transfer of metal ions between an aqueous phase and an organic one involves the formation of small mononuclear clusters typical of metal–ligand coordination chemistry, at one extreme, in the organic phase, and their aggregation to multinuclear primary clusters that self-assemble to form even larger superclusters typical of supramolecular chemistry, at the other. Our metrical results add an orthogonal perspective to the energetics-based view of phase splitting in chemical separations known as the micellar model—founded upon the interpretation of small-angle neutron scattering data—with respect to a more general phase-space (gas–liquid) model of soft matter self-assembly and particle growth. The structure hierarchy observed in the aggregation of our quinary (zirconium nitrate–nitric acid–water–tri-n-butyl phosphate–n-octane) system is relevant to understanding solution phase transitions, in general, and the function of engineered fluids with metalloamphiphiles, in particular, for mass transfer applications, such as demixing in separation and synthesis in catalysis science.

Synopsis

Multiscale structure perspectives for zirconium-loaded organic phases show hierarchical aggregation built upon coordination complexes that self-assemble into primary clusters that form superclusters.

Introduction

ARTICLE SECTIONS
Jump To

The literature of science and engineering of colloids is filled with ternary phase diagrams comprising two immiscible liquids (e.g., water and oil) and a soluble amphiphilic surfactant. (1,2) Upon mixing, the initially biphasic system can equilibrate as one of four main types of microemulsions described by Winsor. (3) In each of these oil-in-water and water-in-oil phase combinations, the micelle (4) and the reverse micelle, (5) respectively, are the fundamental building blocks in the formation of supramolecular solution structures, including cylindrical, close-packed, and lamellar mesophases. (6,7) In the vernacular of soft matter sciences, such complex (structured) fluids are built upon multiscale correlations resulting from physicochemically soft and weak interactions. The organization of solutes (i.e., neutrals, cations, and anions) into aggregated structures through weak interactions, with energies of 1 kBT and less, (8,9) drives the assembly of aggregate architectures in which a delicate balance between short-range attractive, long-range repulsive (SALR) interactions regulates pattern formation (10) and phase behaviors. Whereas the weak long-range forces are small compared the strong short-range forces in metalloamphiphile chemical bonding (approximately 300 kBT), they influence complex physical responses, often resulting in complicated fluid behaviors, including the critical phenomenon of phase splitting.
In the limit of a vanishing critical temperature and weak attraction, the splitting of SALR fluids into colloid-poor (gas) and colloid-rich (liquid) phases—known as the gas–liquid phase separation model (11,12)—is suppressed (10) in favor of percolation. (13,14) In contrast, strong attraction energy landscapes between nanoscale solute architectures leading to gelation in colloidal SALR systems are a direct consequence of equilibrium gas–liquid phase separation, not percolation. (15) Recently, self-assembly phenomena in SALR colloids at macroscopic length scales were attributed to a previously unrecognized mechanism, dubbed ionic sphericity. (16) In previous research on complex fluids that function in ternary (water–oil–surfactant) systems for chemical separations by liquid–liquid (biphasic) extraction, the splitting of the oil phase—into two oil phases—has been attributed to the gas–liquid model of phase transition. (17) This interpretation, originally proposed in 1998 by Erlinger et al. (18) on the basis of small-angle X-ray scattering (SAXS), is the genesis of the micellar model of phase splitting. (19) In this, a complex fluid of sticky spheres (e.g., reverse micelles with surface adhesion) with a short-range interaction potential that is approximated by a hard repulsive core and a rectangular attractive well (as defined by Baxter) (20) was shown to be consistent with the experimental SAXS data. Despite the system approximations and simplifications required to make the data analyses tractable, the insights obtained from the use of this model with regard to the energetics of the phase transition provide a zeroth-order benchmark for all subsequent phenomenological comparisons, including those reported here.
To address aspects of the complexities with regard to organic phase architectures, in general, and the origins of phase splitting, in particular, we provide new insights into the soft matter chemistry and physics of separation science. In particular, lacking in our understanding of these systems is a robust description of the weak interactions involved in aggregation and how they scale up (and down) as manifest in the type of solution structures that are formed, and how, in combination, they influence technically relevant material behaviors. For example, self-organized, complex fluid systems are exemplified by a wide variety of familiar substances, including polymers, gels, soaps, and inks. Recent advances with X-ray and neutron scattering probes focused on fluid systems have significantly expanded the list of liquids exhibiting aggregation behaviors in a diverse range of applications. Some of these include commercial-scale processes for enhanced oil and gas recovery; production of pharmaceuticals and commodity chemicals; transportation and sequestration of natural gas and CO2; phase transfer and micellar catalysis; and even nuclear fuel reprocessing operations like PUREX (plutonium uranium reduction extraction). (21)
In this latter system and many others like it, there is an emerging consensus that solute aggregation can have a profound impact on the chemical behavior of the organic solution, notably on both the efficiency and selectivity of the process. (22−33) In fact, reverse micellar solvents with the macroscopic appearance of a homogeneous solution are routinely employed in organic- and biological-based separation, purification, and conversion processes with little understanding of the morphology and arrangement of the micelle domains. (34,35) These contemporary applications are revolutionizing fundamental separations concepts by shifting the focus away from the classical enthalpic drivers associated with local, molecular-scale interactions into multiscale-ordering, which is hypothesized to include large entropic contributions that may dominate the energy landscape of bulk behaviors (36) that are hallmarks of coacervation (35) and microemulsions. (37) However, the relationships between the microscopic solute structures and phase behaviors remains largely unresolved. The present study investigates the macroscopic splitting of the organic phase during liquid–liquid (water–oil) extraction by examining the microscopic structures of metal coordination chemistry and the mesoscopic architectures of supramolecular chemistry. Our findings provide a fundamental, telescoping structure perspective into processes that are pivotal to the international energy economies vis-à-vis the refining and purification of platinum group and base metals as well as rare earth and actinide elements.
In these biphasic hydrometallurgical operations, a high metal loading of the organic phase is advantageous to obtain cost-effective separations. However, in practice, it is not possible to achieve this because when the concentration of the extracted metal ion exceeds the critical concentration, a ruinous phase transition occurs. In this, the organic phase splits into a light phase and a heavy one; the heavy organic phase is called the third phase. (21,38) In line with recent research on the subject, (39) we choose to cast organic phase splitting in liquid–liquid extraction as a critical phenomenon. Understanding the mechanism of third-phase formation offers the prospects for the development of advanced, high-performance separation science and technology. With specific regard to the PUREX process—on a molecular scale—it is generally known that the amphiphile tri-n-butyl phosphate (TBP; see the Supporting Information showing the chemical structure) coordinates with tetravalent plutonium as the nitrate complex to form the charge-neutral tetranitrato-disolvate, Pu(NO3)4(TBP)2, (38) in the organic phase. On a supramolecular scale, small-angle neutron scattering (SANS) has been modeled in the Baxter sticky-sphere framework to reveal interactions between small spherical reverse micelles containing three to five TBP molecules in n-dodecane. (40) The critical phenomenon of phase splitting was attributed to the strength of the attractive forces (with a potential energy of up to −2.6 kBT) between the polar cores of the reverse micelles. This is dubbed the micellar interaction model. (41) More recent reports involving computational studies have called into question the validity of the model for both metal-free (42−44) and metal-ion-loaded (45) systems.
As with any model, new experimental and computational data obtained with contemporary methodology oftentimes reveal nuances and subtleties that are not resolved in the original efforts. (41) For example, Baldwin et al. (45) reported that the Baxter model for interpreting SANS profiles in the PUREX system can be misleading about the strength of the interaction energies between reverse micelles. That is, under selected conditions, the model tends to predict nonattractive interactions contrary to direct observations of phase splitting due to strong attractive associations. There are at least two possible reasons for these contradictions. First, to simplify analysis with the Baxter model, the extracted solutes were assumed to occupy the cores of spherical or ellipsoidal micelles without knowledge of the true coordination chemistry and particle morphologies. Second, the incoherent scattering intensity from hydrogen atoms contained in the samples was overly simplified as a fitting parameter in SANS data analyses. (41,45−47) Since the requisite coherent scattering intensity reflecting the structural information on the solutes is typically buried under the level of incoherent scattering intensity, it is crucial to evaluate the incoherent scattering intensity in a rigorous manner. Accordingly, in this study we used neutron polarization analysis, which is the best experimental method for evaluating the incoherent scattering intensity to avoid misinterpretation of the SANS data. (48) This development of an improved method to model long-range aggregate structures beyond coordination chemistry augments our understanding of the solution state of the organic phase following extraction.
In this work, extended X-ray adsorption fine structure (EXAFS) experiments with density functional theory (DFT) calculations, molecular dynamics (MD) simulations, and SANS were carried out in a complementary style to explore the micro- and mesoscopic architectures of the solvate cluster formed by Zr(NO3)4 and TBP, Zr(NO3)4(TBP)2, as a nonradioactive chemical surrogate of Pu(NO3)4(TBP)2 in PUREX. The combination of experimentation and simulation is of great advantage to avoid the above-mentioned uncertainties by accessing a broad spectrum of length scales from 0.1 to 100 nm in a telescoping, hierarchical model (49) that includes the extracted coordination complexes (obtained from EXAFS) and also the morphology of aggregates (by SANS). This multimodal method allows unique data treatment, enabling us to combine EXAFS and SANS with theory, (50,51) to elucidate the molecular nature of the zirconium complexes and their subsequent self-assembly into supramolecular aggregates. The results advance our knowledge about the structures and energetics responsible for the phenomena of phase transitions in SALR fluids as they approach critical point concentrations.

Results and Discussion

ARTICLE SECTIONS
Jump To

Experimental Samples

All samples used for EXAFS and SANS measurements were prepared by biphasic solvent extraction of the Zr(NO3)4–HNO3/TBP–n-octane-d18 system. Equal volumes (3.0 mL) of an aqueous, initial phase of 10.5 M (mol/L) nitric acid solution containing [Zr(NO3)4]aq,in = 0 mM and the organic, initial phase of 0.50 M TBP were shaken for 1 h at room temperature in a glass tube, where [Zr(NO3)4]aq,in is the aqueous, initial concentration of Zr(NO3)4 (prior to the extraction). After centrifugation and separation of the two distinct phases, the organic phase was collected; this is designated as sample no. 1. Next, 3.0 mL of sample no. 1 and 3.0 mL of aqueous, initial phases containing [Zr(NO3)4]aq,in ranging from 0.010 to 0.049 M in 10.5 M nitric acid solution were shaken for 1 h at room temperature. The mixtures were centrifuged to separate the aqueous and organic phases. Aliquots of the organic phases are designated as sample nos. 2–5 (Table 1). Sample nos. 1–5 were examined by both EXAFS and SANS. Note that the phase separation and third-phase formation occur only in cases where [Zr(NO3)4]aq,in > 0.049 M. Therefore, the complex concentration in sample no. 5 is very close to the critical concentration, i.e., the highest Zr concentration in the organic phase achievable without phase separation. In this extraction, Zr(NO3)4(TBP)2 is the predominant extracted coordination species in the organic phase, as will be explained in conjunction with analysis of the EXAFS data. The concentrations of Zr(NO3)4(TBP)2, HNO3, and H2O extracted into the organic, equilibrium phases of all samples are denoted by [Zr(NO3)4(TBP)2]org,eq, [HNO3]org,eq, and [H2O]org,eq, respectively. The concentration of free TBP (uncoordinated with Zr(NO3)4) in the organic phases is denoted by [TBP]org,eq. Additionally, the aqueous, equilibrium concentrations of Zr(NO3)4, [Zr(NO3)4]aq,eq, are shown in Table 1. The values of [Zr(NO3)4]aq,in, [Zr(NO3)4(TBP)2]org,eq, [HNO3]org,eq, [H2O]org,eq, and [TBP]org,eq of the five samples are summarized in Table 1, together with the volume fractions of extracted Zr(NO3)4(TBP)2, HNO3, H2O, uncoordinated TBP, and n-octane-d18, which are defined as ϕZr(NO3)4(TBP)2, ϕHNO3, ϕH2O, ϕTBP, and ϕoctane-d18, respectively. Note that ϕZr(NO3)4(TBP)2 + ϕHNO3 + ϕH2O + ϕTBP + ϕoctane-d18 = 1. The distribution ratios of zirconium, DZr, are shown in Table 1. The values (1.9–2.4) are independent of [Zr(NO3)4]aq,eq, indicating that zirconium is extracted into the organic phases as a mononuclear entity. This represents a stark contrast with the distribution ratios of tetravalent cerium, DCe, reported elsewhere, in which cerium reports to the organic phase as a multinuclear complex. (52) The concentration ratios of [HNO3]org,eq/[Zr(NO3)4(TBP)2]org,eq, [H2O]org,eq/[Zr(NO3)4(TBP)2]org,eq, and [HNO3]org,eq/[H2O]org,eq for sample nos. 1–5 are also shown in the Supporting Information.
Table 1. Composition of the Aqueous (aq) and Organic (org) Phases before (in) and after (eq) the Extraction Process and Distribution Ratio of Zirconium Ions, DZr
     [Zr(NO3)4(TBP)2]org,eq (M)[HNO3]org,eq (M)[H2O]org,eq (M)[TBP]org,eq (M) 
sample no.[Zr(NO3)4]aq,in (M)[HNO3]aq,in (M)[Zr(NO3)4]aq,eq (M)DZrϕZr(NO3)4(TBP)2ϕHNO3ϕH2OϕTBPϕoctane-d18
1010.5  00.280.0450.500 
  00.0110.0010.1170.871
20.01010.50.0032.30.0070.280.0510.486 
0.0050.0070.0010.1320.855
30.02510.50.0082.10.0170.340.0640.466 
0.0130.0140.0010.1270.845
40.03410.50.0102.40.0240.270.0760.452 
0.0190.0110.0010.1040.865
50.04910.50.0171.90.0320.300.0740.436 
0.0250.0130.0010.1190.842

Local Coordination Structure in the Organic Phase

Figure 1a,b show the k3-weighted Zr K-edge EXAFS oscillation, k3χ(k), as a function of the photoelectron wavenumber, k–1), and the corresponding radial structural function, |FT[k3χ(k)]|, with the imaginary part of FT[k3χ(k)], Im{FT[k3χ(k)]} obtained for the organic phases (sample nos. 2–5) as a function of the radial distance from the Zr atom, r + Δr. Here, Δr is the magnitude of the phase shift resulting from a change in the photoelectron wave while traversing the potentials of the absorbing and scattering atoms. The k3χ(k) in Figure 1a and |FT[k3χ(k)]| in Figure 1b do not depend on the increase in [Zr(NO3)4(TBP)2]org,eq in the organic phase. Importantly, this indicates that the local coordination structure of the extracted zirconium complex barely changes at [Zr(NO3)4(TBP)2]org,eq ≤ 32 mM, which is close to the highest complex concentration attainable in the organic phase without third-phase formation. Even in the case of the third phase, the EXAFS data were similar to those of the organic phases, as shown in Figure S6. This mirrors the response observed in the corresponding Ce EXAFS data obtained under comparable conditions. (53)

Figure 1

Figure 1. EXAFS spectra for the extracted Zr coordination complexes, (a) k3-weighted Zr K-edge EXAFS, k3χ(k) (open black circles), and (b) corresponding Fourier transform, |FT[k3χ(k)]| (open black circles), and the imaginary part of FT[k3χ(k)], Im{FT[k3χ(k)]} (filled blue circles), obtained for the organic phases of sample nos. 2–5. The solid black curves in part a, the red curves in part b, and the green curves in part b are the simulated k3χ(k), |FT[k3χ(k)]|, and Im{FT[k3χ(k)]} responses, respectively. The thick arrows highlight the P1, P2, P3, and P4 peaks, which originate from the scattering paths of Zr–OTBP and Zr–ONO3, Zr–NNO3, Zr–PTBP, and Zr–N–Omultiple, respectively (Table 2).

In Figure 1b, the |FT[k3χ(k)]| shows one significant peak with a shoulder on the longer r + Δr side. The peak, at approximately r + Δr ≈ 0.15 nm and designated as P1, is indicated with a thick arrow. Three further peaks are designated as P2, P3, and P4 at 0.2 nm < r + Δr < 0.4 nm. Peak P1 is attributed to two kinds of Zr–O bonds, namely, directly coordinating oxygen atoms of TBPs and nitrate ions. The peaks P2, P3, and P4 are attributed to the correlations between Zr–P with TBPs, Zr–N with nitrate ions, and/or multiple scattering of Zr–N–O with nitrate ions, respectively.
Curve fitting for quantitative analysis of the EXAFS data was conducted to obtain the structural parameters of the extracted coordination species in the organic phases. In this analysis, the atomic positions from an optimized coordination structure of Zr(NO3)4(TBP)2 obtained by DFT calculation were used as a structural model for initial input in simulation code FEFF version 8.4. (54) The optimized coordination structure is shown in Figure 2a. All significant scattering paths between Zr and its interrelated atoms giving rise to the distinct peaks in |FT[k3χ(k)]| were taken into account in the curve-fitting analysis. The amplitude reduction factor, S02, was fixed as 0.9 in the fitting procedure. All EXAFS data were well reproduced by the best-fitted simulated curves of k3χ(k) (black solid lines), |FT[k3χ(k)]| (red solid lines), and Im{FT[k3χ(k)]} (green solid lines) with the refined structural parameters, as shown in Figure 1.

Figure 2

Figure 2. Schematic diagrams of hierarchical aggregate model of zirconium superclusters. (a) Geometry of the optimized coordination structure of extracted Zr(NO3)4(TBP)2 in the organic phase, determined by DFT calculation. Green, Zr; yellow, P; red, O; blue, N; black, C; and light pink, H. (b) Primary cluster in which the Zr(NO3)4(TBP)2 complexes (red spheres) distribute with radius RS around the central complex, (c) primary clusters assemble into a large aggregate (supercluster), where the primary clusters with radius RS surround the central cluster (light blue sphere) with radius RL. A set of the number of the primary clusters, M = 25, and the number of the complexes, N = 7, corresponds to the characteristic parameters of sample no. 5 from SANS data analysis.

Representative structural parameters, refined in the fitting analysis of sample no. 4 ([Zr(NO3)4(TBP)2]org,eq = 24 mM), are summarized in Table 2, where CN, rEXAFS, σDB2, and ΔE0 correspond to the coordination number, bond distance, squared Debye–Waller factor, and energy shift parameter, respectively. The oxygen CN due to TBP molecules (Zr–OTBP path) and nitrate ions (Zr–ONO3 path), resulting in P1, are approximately 2 and 8, respectively. The nitrogen CN due to nitrate ions (Zr–NNO3 path), phosphorus due to TBPs (Zr–PTBP path), and multiple scattering due to nitrate ions (Zr–N–Omultiple path), resulting in P2, P3, and P4, are approximately 4, 2, and 4, respectively. Accordingly, the CNs of the Zr–OTBP and Zr–PTBP paths indicate the coordination of two TBP molecules with the inner sphere of zirconium. The O and N CNs of Zr–ONO3, Zr–NNO3, and the Zr–N–Omultiple paths indicate the bidentate coordination of four nitrate ions with the inner sphere of zirconium. As a result, we found that the coordination species in the organic phase is always Zr(NO3)4(TBP)2. This species is therefore expected to be a fundamental building unit of a higher-order, long-range structure formed in the organic phases.
Table 2. Structural Parameters Evaluated by Fitting Analysis of Sample No. 4 Using an Optimized Coordination Structure of Zr(NO3)4(TBP)2 as the Initial Input Model for DFT Calculation
pathaCNrEXAFS (nm)σDW2b (nm2)ΔE0b (eV)S02rDFT (nm)
Zr–OTBP1.9 ± 0.21c0.215 ± 0.0020.000 023.630.90.2205
Zr–ONO37.8 ± 0.370.228 ± 0.0030.000 073.630.90.2349
Zr–NNO33.9 ± 0.330.276 ± 0.0040.000 033.630.90.2795
Zr–PTBP2.1 ± 0.200.361 ± 0.0030.000 043.630.90.3655
Zr–N–Omultiple4.1 ± 0.320.398 ± 0.0030.000 033.630.90.4003
a

Scattering paths calculated by use of program code FEFF version 8.4. (54)

b

Errors in σDW2 and ΔE0 in this study are within ±3.0% accuracy.

c

Error represents ±1 standard deviation throughout the paper.

The result of the DFT calculation shows that directly coordinated oxygen atom of only one nitrate ion is slightly more distant than those of the other three nitrate ions coordinated to Zr, thereby giving rise to the shoulder on peak P1 and to the increase of σDW2 of Zr–ONO3. The average bond distances evaluated by DFT calculation, rDFT, are shown in Table 2 and are in good agreement with rEXAFS. Therefore, the optimized coordination structure of Zr(NO3)4(TBP)2 shown in Figure 2a can be taken as the actual state in the organic phase.

Overall SANS Features

Figure 3a shows the observed SANS intensity distribution as a function of q, Iobs(q), obtained for sample nos. 1–5, which had different concentrations of Zr(NO3)4(TBP)2 in their organic phases, as shown in Table 1. Here, q (=[4π/λ]sinθ) is the magnitude of the scattering vector, and λ and 2θ are the wavelength of the incident neutrons and the scattering angle, respectively. The scattering intensity in the low-q region (q < 1.0 nm–1) gradually increases with increasing [Zr(NO3)4(TBP)2]org,eq, whereas that in the high-q region (q > 1.0 nm–1) barely changes as a function of [Zr(NO3)4(TBP)2]org,eq. Even in the absence of extracted Zr(NO3)4(TBP)2 complex in the organic phase (sample no. 1), scattering intensity is observed in the SANS profile, particularly in the low-q region. This is attributed to aggregates formed by extracted HNO3 and TBP, which have been reported by SANS (41) and computational studies. (43−45) Our preceding research concluded that the TBP aggregates arise from the formation of hydrogen-bonding networks consisting of HNO3, H2O, and the P═O group of TBP, leading to an isotropic microemulsion in the organic phase. (44) Because the scattering from the HNO3–H2O–TBP aggregates for sample nos. 2–5 can vitiate the quantitative structural analyses of the extracted Zr(NO3)4(TBP)2 in the organic phase, the undesirable scattering contributions were subtracted deliberately as described in the Methods section (see the Supporting Information).

Figure 3

Figure 3. Double-logarithmic plots of the SANS profiles, (a) Iobs(q) and (b) Isub(q), with error bars, as a function of [Zr(NO3)4(TBP)2]org,eq. [Zr(NO3)4(TBP)2]org,eq gradually increases with sample no. from 1 to 5. Dashed lines in part b are the form factors of Zr(NO3)4(TBP)2 in the organic phase, determined on the basis of the Debye scattering formula for randomly orientated Zr(NO3)4(TBP)2 using eqs S3, S6, and S7. Solid lines in part b are the best-fit theoretical SANS profiles obtained by using eqs S10–S13 together with the characteristic parameters listed in Table 3.

Figure 3b shows Isub(q) plots, which are the SANS profiles of sample nos. 2–5 from which the components of the HNO3–H2O–TBP aggregates were subtracted from Iobs(q). The scattering intensity in the high-q region (q ≥ 3.0 nm–1) increases in proportion to the increase in [Zr(NO3)4(TBP)2]org,eq, indicating that the coordination structure of the extracted metal ion species does not depend on the concentration and maintains a specific structure, which is consistent with the EXAFS results discussed above. In addition, the scattering intensity in the low-q region (q < 1.0 nm–1) increases out of proportion to [Zr(NO3)4(TBP)2]org,eq suggesting that Zr(NO3)4(TBP)2 does not disperse homogeneously and, rather, forms supramolecular structures.

Quantitative Analyses of SANS Intensity Distributions

We first numerically analyzed the form factor of a single Zr(NO3)4(TBP)2 complex in sample nos. 2–5 using the optimized coordination of each atom with the DFT calculation by means of the Debye scattering formula for randomly orientated objects; (50,51) the scattering function is described in detail in the Supporting Information (see eqs S3−S9). The four dashed lines in Figure 3b indicate the same q dependence and show the contribution of the form factors of Zr(NO3)4(TBP)2 for sample nos. 2–5. The small-angle scattering intensities are proportional to the number density of the extracted Zr(NO3)4(TBP)2 complex, ncomplex, and agree well with Isub(q) for q ≥ 3.0 nm–1, whereas those for q < 3.0 nm–1 deviate from each form factor. This deviation is attributed to an ordered (aggregate) structure well beyond the inner-sphere coordination about Zr, and thus analyzing the excess scattering components for q < 3.0 nm–1 over the form factor of the coordination species will allow elucidation of the aggregate architecture. Analyses of Isub(q) with Guinier’s law (55) as shown in Figure S7 provide an initial evaluation of the size of the high-order structures formed by Zr(NO3)4(TBP)2 in the organic phase in terms of the radius of gyration, Rg. These were determined to be 3.2, 3.3, 4.5, and 5.0 nm for sample nos. 2–5, respectively. The observed power law scattering, Isub(q) ∼ qD for 1/Rg < q (nm–1) < 3, shows D ≈ 3 for sample nos. 3–5, which is solid evidence of the existence of an inhomogeneous spatial distribution of Zr(NO3)4(TBP)2 complexes inside the high-order structures.
Given the above findings, we assumed that the hierarchical structure, which comprises primary clusters consisting of extracted Zr(NO3)4(TBP)2 and its aggregates, is formed in the organic phases as shown in Figure 2. A small cluster of Zr(NO3)4(TBP)2 complexes is considered; that is, the N complexes distribute spherically and randomly around the center complex at a distance R (see Figure 2a,b). Then, the small primary clusters consisting of N complexes with radius RS form a large aggregate (supercluster) of M primary clusters with radius RL (see Figure 2c). A similar structure hierarchy was proposed to explain the pseudoliquid phase behaviors of hetero-polyanion cluster compounds. (56) Based on this hierarchical aggregate model, (49) the small-angle scattering intensity can be exactly calculated as described in the Supporting Information, and the SANS data were compared with the model.
Sample no. 2 has the lowest concentration of extracted Zr(NO3)4(TBP)2 in the organic phase; consequently, Isub(q) can be reproduced by the contribution of the primary cluster without considering the supercluster, and thus M was fixed as 1 (the model shown in Figure 2b). By comparison, a clustering of the primary clusters is crucial to reproducing Isub(q) when [Zr(NO3)4(TBP)2]org ≥ 17 mM (sample nos. 3–5), otherwise the theoretical values based on the model in Figure 2b do not agree with the experimental data as shown in Figures S9 and S10.
The solid lines in Figure 3b show the best-fit theoretical scattering curves for the hierarchical aggregates, which exhibit good agreement between the model and the experiments. The refined characteristic parameters are summarized in Table 3. Note that the number density of the superclusters, nsuper, satisfies the relation ncomplex = nsuperMN, and the distributions of RS and RL were necessary to reproduce the scattering intensity distribution precisely; thus the Schultz distribution with the corresponding standard deviations, σL and σS, respectively, was used. (57) Remarkably, the parameters N and RS relating to the primary cluster barely change in sample nos. 3–5, even with increasing M and RL, which relate to the superclusters. This result suggests that the size and aggregation number of the primary cluster are insensitive to [Zr(NO3)4(TBP)2]org,eq for it to be stable in the organic phase when the fundamental building unit is Zr(NO3)4(TBP)2. We believe that aggregation of the hydrophilic parts of Zr(NO3)4(TBP)2 provides the driving force to form the primary cluster, and that this is driven by coordinated nitrate ions, which form a hydrogen-bonding network involving extracted HNO3 and H2O. (43,44,58) Consequently, the surface of the primary cluster should be stably covered with the hydrophobic butyl groups of TBP in the organic phase as will be explained by the MD simulations described below. Subsequently, we speculate that the process of primary cluster aggregation is similar to the formation of concentrated reverse micelle systems, as supported by previous studies. (41,59,60)
Table 3. Summary of the Characteristic Parameters of Hierarchical Aggregates Determined by Model Analysis of Isub(q)
sample no.NRS (nm)σS (nm)MRL (nm)σL (nm)
29.8 ± 2.41.3 ± 0.140.70 ± 0.151  
37.0 ± 1.30.93 ± 0.020.10 ± 0.059.7 ± 1.62.7 ± 0.012.2 ± 0.04
46.9 ± 1.20.95 ± 0.020.11 ± 0.0519 ± 3.33.2 ± 0.022.3 ± 0.05
57.0 ± 1.00.92 ± 0.030.10 ± 0.0625 ± 3.43.7 ± 0.022.5 ± 0.04

MD Simulations and the Structure of the Primary Clusters

MD simulations were carried out for each of the systems given in Table 1. Trajectories from the simulations indicated that the Zr complexes tended to aggregate to form clusters, namely, the primary clusters described above. To analyze this clustering quantitatively, we calculated the Zr–Zr radial distribution function, g(rZr–Zr), and the corresponding coordination number of Zr with the other Zr, CN(rZr–Zr), shown for system no. 5 in Figure 4a. The g(rZr–Zr) function consists of a large peak followed by an extended tail. This large peak, which corresponds to the closest Zr–Zr distance on average, reaches its maximum at approximately 0.88 nm. This value is close to the RS value of 0.92 ± 0.03 nm in the characteristic parameters of the hierarchical aggregates model. To define a cluster, we used the algorithm of Sevick, (61) with the criterion that two Zr complexes are connected if the Zr–Zr distance is less than 1.2 nm. This distance is approximately where the first peak of the radial distribution function ends. This choice is somewhat arbitrary, but the resulting analysis is in reasonable agreement with the conclusions drawn from observing many snapshots. The distribution of the primary cluster aggregation numbers is shown in Figure 4b for system no. 5. It is clear that primary clusters form, several with aggregation numbers of 5–7, in agreement with the SANS analysis. The shapes of the primary clusters, according to our simulations, are not uniform, as they vary from spherical to elongated aggregations. This observation is supported by the coordination number analysis, CN(rZr–Zr), given in Figure 4a, where CN(rZr–Zr) gives the number of Zr atoms within a distance rZr–Zr of a central Zr atom. The value of CN(rZr–Zr) at rZr–Zr = 1.2 nm, the position of the first minimum of the radial distribution function, is 2.72, indicating that there are, on average, about 2.72 other Zr ions surrounding one Zr ion within a 1.2 nm cutoff range. Smaller clusters are also seen, but it is noteworthy that clusters of three or four Zr atoms are relatively uncommon. To check that this was not an equilibration issue, we ran 10 independent MD simulations, starting from random initial configurations. All the simulations used the same simulation protocols as described in the Supporting Information (Computational Methods). The results, which are shown in Figure S2 in the Supporting Information, indicate that the cluster size distribution is identical, to within statistical uncertainty, for all 10 systems.

Figure 4

Figure 4. (a) Zr–Zr radial distribution function, g(rZr–Zr) (solid line), and corresponding coordination number of Zr with the other Zr, CN(rZr–Zr) (dashed line). (b) The primary cluster probability distribution for system no. 5 as determined by the MD. The ordinate gives the probability of finding a primary cluster with a given number of Zr atoms. This probability is the number of primary clusters of a given aggregation number divided by the total number of the primary cluster.

The system sizes are too small to observe and analyze hierarchical clustering of these primary clusters, but the long broad tail in the radial distribution function in Figure 4a indicates that the primary clusters in the simulated systems are not homogeneously distributed. An examination of simulation snapshots shows the loose aggregation of a small number of primary clusters. An example is given in Figure 5a, where two primary clusters are seen in close proximity, providing an indication of incipient hierarchical structuring.

Figure 5

Figure 5. (a) Snapshot of two, neighboring primary clusters from our MD simulations and (b) magnified snapshot showing the hydrogen-bonding network within the primary clusters. Green, Zr; yellow, P; red, O; blue, N; black, C; light pink, H; and light blue dashed line, hydrogen bond.

We now consider the structure of the primary clusters. Examples are shown in Figure 5b. Within a primary cluster there are hydrogen bonds, involving uncoordinated TBP, HNO3, and H2O. These play a role in binding the Zr complexes together. There are, in addition, nonspecific polar interactions, for example, involving the charged species within the Zr complexes. Hydrophobic interactions between the nonpolar groups are also in evidence. Turning now to the interactions between primary clusters, any inferences based on simulation results must be extremely tentative, due to the system size issues noted previously. That said, the indications are that there is no direct hydrogen bonding between primary clusters and no specific bridging molecules. This is exemplified by Figure 5a. The available simulation data suggest, instead, that primary clusters interact via a combination of nonspecific electrostatic interactions between the charged groups and hydrophobic interactions between the aliphatic groups. It is noteworthy that recent research by Servis et al., (62) on the structure of the aggregates formed by uranyl nitrate and TBP, also indicated that such electrostatic interactions play a key role in stabilizing the clusters.
Finally, we return to the jagged nature of the cluster distribution (Figure 4b). Our analysis shows that this is not related to the choice of the distance criterion for defining a cluster. The results are insensitive to this. As discussed previously, the fact that 10 independent runs give the same distribution leads us to believe this is an equilibrium distribution. To gain insight into what factors might lead to this distribution, we carried out a hydrogen-bonding analysis, counting the number of hydrogen bonds between NO3, HNO3, and H2O. The criterion for determining the presence of a hydrogen bond is that the donor–acceptor distance should be no more than 0.35 nm, and the acceptor–donor–hydrogen angle should be no more than 30°. The oxygen atoms that have covalent bonds with the hydrogen atoms in the H2O and HNO3 molecules were regarded as potential donors, and the electronegative atoms that possess a lone electron pair were regarded as potential acceptors. Stable primary clusters, defined for these purposes as clusters with a lifetime greater than 1 ns, were identified and indexed for hydrogen bonding. These primary clusters contained 1, 2, 3, 5, and 6 Zr complexes. Tetramers are apparently too unstable to survive for this length of time. The average number of hydrogen bonds per Zr atom is plotted against cluster size in Figure 6a.

Figure 6

Figure 6. (a) Number of hydrogen bonds per Zr complex as a function of aggregation number. (b) Radial distribution function for octane carbon atoms around a central Zr atom as a function of cluster aggregation number of Zr complexes per primary cluster. (c) Radial distribution function for octane carbon atoms around a central N atom in a nitrate ligand, shown as a function of cluster aggregation number. All results in parts a–c are obtained for system no. 5.

While the number of hydrogen bonds increases when two monomers form a dimer, there is no increase in the number of hydrogen bonds on forming higher-order clusters. It would thus appear that hydrogen bonding is probably not a governing factor in determining the number of Zr complexes in a primary cluster.
Another factor that might influence the aggregation number is the degree to which the octane solvent is shielded from polar groups in the clusters. Such interactions would be hydrophobically unfavorable. We therefore calculated the radial distribution functions for octane carbon atoms surrounding a central Zr atom and for octane carbon atoms surrounding the N atom of a nitrate ligand. The plots are shown in Figure 6b,c. These two plots show that there is considerably more unfavorable polar–nonpolar contact for the trimer, as compared to either the dimer or the hexamer. Thus, it would appear that the polar groups in the trimer are less well shielded from the diluent than is the case for either the dimer or the hexamer, and this may partly explain the relative stabilities of these clusters.

Thermodynamic Considerations

We relate the extracted parameters from the model fittings of Isub(q) in SANS to the thermodynamic properties of the system. We start with the van’t Hoff equation given by (63)
(1)
where Π is osmotic pressure, v is the occupied volume of the solute per molecule, kB is the Boltzmann constant, and T is the thermodynamic temperature. The osmotic pressure can be related to the forward scattering intensity, Isub(q = 0), by (64)
(2)
with the difference between the scattering length densities of the Zr(NO3)4(TBP)2 complex and the matrix, Δρ, and ϕZr(NO3)4(TBP)2. In the studied system, the volume v in eq 1 can be assumed to be the occupied volume per supercluster, and hence, v ≈ 1/nsuper. The calculated v and the osmotic pressure obtained by eq 1, Πvan’t-Hoff, are listed in Table 4. Because the vapor pressure of octane at 20 °C is 1.33 kPa, and the matrix consists of a mixture of extracted HNO3, extracted H2O, uncoordinated TBP, and n-octane-d18 (Table 1), the calculated Πvan’t-Hoff is a good estimation. The osmotic pressures can also be calculated with eq 2 by using Isub(q = 0) from the experimental data, which were estimated by Guinier plots (Table 4). The osmotic pressures from Isub(q = 0), ΠI(0), and Πvan’t-Hoff agree within 10%, which indicates that our structural model is correct.
Table 4. Thermodynamic Parameter Values
sample no.v (nm3)Isub(q = 0) (cm–1)Πvan’t-Hoff (Pa)ΠI(0) (Pa)RHS (nm)
22.5 × 1030.1221.6 × 1031.8 × 1031.3
36.8 × 1032.436.1 × 1026.0 × 1022.7
48.9 × 1036.724.6 × 1024.9 × 1023.2
59.1 × 10312.44.5 × 1024.4 × 1023.7
The van der Waals equation of state with hard spheres with radius RHS modifies eq 1 to give (63)
(3)
where β = 4ω with ω = 4πRHS3/3 and α = 27kBTcβ/8 with the critical temperature, Tc, for phase separation. By assuming Tc/T = 1, RHS for each sample can be calculated with assigning ΠI(0) and v, which are shown in Table 4. The RHS values agree well with the RL values. The parameter RHS is slightly affected by Tc/T around Tc/T = 1, although the variation of RHS with a ±10% variation of Tc/T = 1 is less than 5%. This result suggests that RL obtained by the model is reasonable.
It is worth noting, at this point, some curious features of this aggregation process. First, as shown in Table 1, the molar concentrations of zirconium in the organic phase are relatively small. The highest concentration studied was sample no. 5, where [Zr(NO3)4(TBP)2]org,eq = 0.032 M, and as noted previously, third-phase formation sets in at concentrations just a little greater than this. A normal solution at such low concentrations might be expected to be approximately ideal-dilute, with Henry’s law applying, but this is clearly far from the case here. To account for the clustering and for the phase transition, it must be the case that the mixture is very far from ideal, implying remarkably strong interactions between the zirconium complexes and, indeed, between the primary clusters. As shown in Table 3, the aggregation number of the primary clusters is approximately 7 and does not change with increasing [Zr(NO3)4(TBP)2]org,eq. This is reminiscent of spherical micelles formed in lyotropic systems, where again there is a favored aggregation number, but it is not clear why this should apply here, where simulation suggests the primary clusters to be far from nicely organized spheres. This is certainly worthy of future study.
Even though the concentrations of primary clusters are small, the results of Table 3 show that they aggregate, indicating strong interactions between them. One may note that the primary clusters, with solvating water, nitric acid, and uncoordinated TBP molecules, are polar entities immersed in a nonpolar, low dielectric medium. This may mean that Coulomb interactions between primary clusters are strong, contributing considerably to the attractive forces between them. Once these superclusters have formed, their concentration is very small, and as shown above, the osmotic pressure may be well estimated by assuming one has an ideal suspension of superclusters.
For values of [Zr(NO3)4(TBP)2]org,eq a little greater than that of sample no. 5, the system separates into two phases. One phase is diluent-rich with little Zr, while the coexisting phase, the third phase, is rich in Zr and has a reduced concentration of diluent. Our data indicate that the aggregation number, M, of the superclusters rises rapidly with [Zr(NO3)4(TBP)2]org,eq, presaging an oncoming phase instability as more and more primary clusters condense. Unfortunately, these data alone are insufficient to cast light on the nature of this phase transition, for this would require information about the structure of the third phase. In the literature, this third phase transition has been treated both as a gas–liquid transition and as a reverse micelle to microemulsion or gel transition. Taken literally, (65,66) the gas–liquid transition model would correspond to a gas of primary clusters (or reverse micelles) condensing to a liquid of primary clusters, but with the primary clusters retaining their integrities. In the second scenario, the primary clusters/reverse micelles would lose their individual identities and merge to form a connected, bicontinuous structure. Of course, another possibility is that the third phase is a bicontinuous emulsion of polar and nonpolar regions, such as appears to be the case in metal-free systems, (44) with distinct Zr clusters embedded in the polar regions of this emulsion. Further study is needed to clarify this situation, but what we have shown is that the strong attractions between primary Zr clusters are the driving force that makes the “gas” phase unstable.

Conclusion

ARTICLE SECTIONS
Jump To

We have clarified the microscopic structure of the organic phases containing extracted Zr(NO3)4(TBP)2 using a method combining EXAFS with DFT calculations, MD simulations, and SANS observations. The coordination structure of the extracted Zr(NO3)4(TBP)2 in the organic phase does not depend on the concentration of Zr(NO3)4(TBP)2 and maintains a specific structure as a fundamental building unit to form high-order (aggregate) architectures. Moreover, quantitative analyses of SANS data using an accurate evaluation of the incoherent scattering intensity and a form factor of randomly oriented Zr(NO3)4(TBP)2 provide a clearer picture of the high-order structures on the basis of scattering theory. Zr(NO3)4(TBP)2 was found to form a hierarchical aggregate composed of small primary clusters comprising a supercluster (Figures 2 and 5). A hybrid interaction consisting of a hydrogen-bonding network for the primary clusters and an attractive interaction for the superclusters induces the formation of the hierarchical aggregate. An increase in [Zr(NO3)4(TBP)2]org,eq has little effect on the size of the primary cluster but increases the size of the superclusters. Furthermore, MD simulations provided direct evidence about the formation of primary clusters, which are assembled by the hydrogen-bonding network involving uncoordinated TBP, HNO3, and H2O. In the clustering analyses of the MD snapshots, the distribution of the primary cluster aggregation numbers of 5–7 agreed well with the SANS data analysis. Accordingly, we conclude that growth of the superclusters is due to an increase in the number of small primary clusters, causing third-phase formation. In this paper, we quantitatively applied the simple aggregation model to analyze SANS profiles. The number of the clusters and their sizes, obtained by SANS data analysis, are fully compliant with van der Waals equations of state. These findings suggest that such behaviors are general ones, extending beyond zirconium in the PUREX process to a wide variety of extraction systems (e.g., DIAMEX, TRUEX, ALSEP) (38) that show multiscale ordering. There is then the intriguing possibility that, by tailoring the interactions between these superclusters, we may afford a new entry into the design of next-generation ionic separation techniques with complex fluids typical of soft matter chemistry.
It is known that the addition of phase modifiers, such as aliphatic alcohols, into organic phases during liquid–liquid extraction allows a higher loading of metal ions (without third-phase formation) than in unmodified systems. (67,68) The fact that the separation selectivity and efficiency are also impacted leads us to speculate that modifiers contribute to the formation of hydrogen-bonding networks inside the primary clusters, and these networks vary in accordance with the intra- and intercluster interactions. In view of the largely empirical use of modifiers in practical chemical separations, the combination of SANS, EXAFS, and computational methods used here can help to elucidate the role of the modifiers in other solvent extraction systems, contributing to the development of high-performance processes in chemical separation science.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acscentsci.8b00669.

  • Method section and additional data and figures including cluster probabilities, configurations, radial distribution function, chemical structure, EXAFS data, Guinier plots, theoretical modeling of the hierarchical aggregates for SANS data analysis, simulated SANS profiles, and coordination structure of zirconium nitrate complex in initial aqueous phase (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Authors
  • Authors
    • Tohru Kobayashi - Materials Sciences Research Center, Japan Atomic Energy Agency, Tokai, Ibaraki 319-1195, Japan
    • Hitoshi Endo - Materials Sciences Research Center, Japan Atomic Energy Agency, Tokai, Ibaraki 319-1195, JapanNeutron Science Division, Institute of Materials Structure Science, and Materials and Life Science Division, J-PARC Center, High Energy Accelerator Research Organization, 203-1 Shirakata, Tokai, Ibaraki 319-1106, JapanDepartment of Materials Structure Science, The Graduate University for Advanced Studies (SOKENDAI), 203-1 Shirakata, Tokai, Ibaraki 319-1106, Japan
    • Junju Mu - School of Chemical Engineering and Analytical Science, The University of Manchester, Oxford Road, Manchester M13 9PL, United KingdomOrcidhttp://orcid.org/0000-0002-7382-2182
    • Christopher D. Williams - School of Chemical Engineering and Analytical Science, The University of Manchester, Oxford Road, Manchester M13 9PL, United KingdomOrcidhttp://orcid.org/0000-0002-5073-5924
    • Mark R. Antonio - Chemical Sciences & Engineering Division, Argonne National Laboratory, Lemont, Illinois 60439, United StatesOrcidhttp://orcid.org/0000-0002-1208-4534
    • William T. Heller - Neutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United StatesOrcidhttp://orcid.org/0000-0001-6456-2975
    • Michihiro Nagao - NIST Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg, Maryland 20899-6102, United StatesCenter for Exploration of Energy and Matter, Department of Physics, Indiana University, Bloomington, Indiana 47408, United StatesOrcidhttp://orcid.org/0000-0003-3617-251X
  • Notes
    The authors declare no competing financial interest.

    Safety statement: no unexpected or unusually high safety hazards were encountered during the course of the experiments.

Acknowledgments

ARTICLE SECTIONS
Jump To

This work was supported in part by the Ministry of Education, Culture, Sports, Science and Technology, Japan (Grant-in-Aid for Scientific Research B, 2014–2018, 26289368, and 2018–2022, 18H01921). H.E. and R.M. thank Prof. Tsuyoshi Koga of Kyoto University for helpful discussions about scattering theory. Part of this research used resources at the Spallation Neutron Source, a DOE Office of Science User Facility operated by Oak Ridge National Laboratory. The synchrotron radiation experiment at the BL11XU beamline of SPring-8 was performed with the approval of the Japan Atomic Energy Agency (JAEA; Proposal 2013A3504). We thank Dr. Shinichi Suzuki and Dr. Tsuyoshi Yaita for discussions and Dr. Hideaki Shiwaku for generous technical support with the EXAFS experiments at SPring-8. Access to the NG5-NSE was provided by the Center for High Resolution Neutron Scattering, a partnership between the National Institute of Standards and Technology (NIST) and the National Science Foundation under agreement DMR-1508249. M.N. acknowledges funding support of cooperative agreement 70NANB15H259 from NIST, U.S. Department of Commerce. The identification of any commercial product or trade name does not imply endorsement or recommendation by the NIST. M.A. acknowledges the support of the U.S. Department of Energy (DOE), Office of Science, Office of Basic Energy Sciences, Chemical Sciences, Geosciences, and Biosciences Division, under contract DE-AC02-06CH11357.

References

ARTICLE SECTIONS
Jump To

This article references 68 other publications.

  1. 1
    Gompper, G.; Schick, M. In Phase Transitions and Critical Phenomena; Domb, C., Lebowitz, J., Eds.; Academic Press: London, 1994; Vol. 16, pp 1176.
  2. 2
    Schwuger, M. J.; Stickdorn, K.; Schomacker, R. Microemulsions in technical processes. Chem. Rev. 1995, 95, 849864,  DOI: 10.1021/cr00036a003
  3. 3
    Winsor, P. A. Solvent Properties of Amphiphilic Compounds; Butterworths: London, 1954.
  4. 4
    Lindman, B.; Wennerström, H. Miceles. Amphiphile aggregation in aqueous solution. Top. Curr. Chem. 1980, 87, 183,  DOI: 10.1007/BFb0048488
  5. 5
    Eicke, H.-F. Surfactants in nonpolar solvents. Aggregation and micellization. Top. Curr. Chem. 1980, 87, 85145,  DOI: 10.1007/BFb0048489
  6. 6
    Fischer, S.; Exner, A.; Zielske, K.; Perlich, J.; Deloudi, S.; Steurer, W.; Lindner, P.; Forster, S. Colloidal quasicrystals with 12-fold and 18-fold diffraction symmetry. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 18101814,  DOI: 10.1073/pnas.1008695108
  7. 7
    Zemb, T. Flexibility, persistence length and bicontinuous microstructures in microemulsions. C. R. Chim. 2009, 12, 218224,  DOI: 10.1016/j.crci.2008.10.008
  8. 8
    Zemb, T.; Holmberg, K.; Kunz, W. Special topic: Weak self assembly. Curr. Opin. Colloid Interface Sci. 2016, 22, A1A3,  DOI: 10.1016/j.cocis.2016.04.001
  9. 9
    Zemb, T.; Kunz, W. Weak aggregation: State of the art, expectations and open questions. Curr. Opin. Colloid Interface Sci. 2016, 22, 113119,  DOI: 10.1016/j.cocis.2016.04.002
  10. 10
    Pini, D.; Parola, A. Pattern formation and self-assembly driven by competing interactions. Soft Matter 2017, 13, 92599272,  DOI: 10.1039/C7SM02125A
  11. 11
    Foffi, G.; De Michele, C.; Sciortino, F.; Tartaglia, P. Scaling of dynamics with the range of interaction in short-range attractive colloids. Phys. Rev. Lett. 2005, 94, 078301,  DOI: 10.1103/PhysRevLett.94.078301
  12. 12
    Zaccarelli, E. Colloidal gels: equilibrium and non-equilibrium routes. J. Phys.: Condens. Matter 2007, 19, 323101,  DOI: 10.1088/0953-8984/19/32/323101
  13. 13
    Bera, M. K.; Qiao, B. F.; Seifert, S.; Burton-Pye, B. P.; Olvela de la Cruz, M.; Antonio, M. R. Aggregation of heteropolyanions in aqueous solutions exhibiting short-range attractions and long-range repulsions. J. Phys. Chem. C 2016, 120, 13171327,  DOI: 10.1021/acs.jpcc.5b10609
  14. 14
    Poon, W. C. K.; Haw, M. D. Mesoscopic structure formation in colloidal aggregation and gelation. Adv. Colloid Interface Sci. 1997, 73, 71126,  DOI: 10.1016/S0001-8686(97)90003-8
  15. 15
    Lu, P. J.; Zaccarelli, E.; Ciulla, F.; Schofield, A. B.; Sciortino, F.; Weitz, D. A. Gelation of particles with short-range attraction. Nature 2008, 453, 499505,  DOI: 10.1038/nature06931
  16. 16
    Kim, S. A.; Jeong, K. J.; Yethiraj, A.; Mahanthappa, M. K. Low-symmetry sphere packings of simple surfactant micelles induced by ionic sphericity. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 40724077,  DOI: 10.1073/pnas.1701608114
  17. 17
    Bauer, C.; Bauduin, P.; Dufreche, J. F.; Zemb, T.; Diat, O. Liquid/liquid metal extraction: Phase diagram topology resulting from molecular interactions between extractant, ion, oil and water. Eur. Phys. J.: Spec. Top. 2012, 213, 225241,  DOI: 10.1140/epjst/e2012-01673-4
  18. 18
    Erlinger, C.; Gazeau, D.; Zemb, T.; Madic, C.; Lefrancois, L.; Hebrant, M.; Tondre, C. Effect of nitric acid extraction on phase behavior, microstructure and interactions between primary aggregates in the system dimethyldibutyltetradecylmalonamide (DMDBTDMA)/n-dodecane/water: A phase analysis and small angle X-ray scattering (SAXS) characterisation study. Solvent Extr. Ion Exch. 1998, 16, 707738,  DOI: 10.1080/07366299808934549
  19. 19
    Testard, F.; Zemb, T.; Bauduin, P.; Berthon, L. In Ion Exchange and Solvent Extraction: A Series of Advances; Moyer, B. A., Ed.; CRC Press: Boca Raton, FL, 2010; Vol. 19, pp 381428.
  20. 20
    Baxter, R. J. Percus-Yevick equation for hard spheres with surface adhesion. J. Chem. Phys. 1968, 49, 27702774,  DOI: 10.1063/1.1670482
  21. 21
    Schulz, W. W.; Navratil, J. D. Science and Technology of Tributyl Phosphate; CRC Press: Boca Raton, FL, 1984.
  22. 22
    Osseo-Asare, K. Aggregation, reversed micelles, and microemulsions in liquid-liquid extraction: the tri-n-butyl phosphate-diluent-water-electrolyte system. Adv. Colloid Interface Sci. 1991, 37, 123173,  DOI: 10.1016/0001-8686(91)80041-H
  23. 23
    Diss, R.; Wipff, G. Lanthanide cation extraction by malonamide ligands: from liquid-liquid interfaces to microemulsions. A molecular dynamics study. Phys. Chem. Chem. Phys. 2005, 7, 264272,  DOI: 10.1039/B410137E
  24. 24
    Jensen, M. P.; Yaita, T.; Chiarizia, R. Reverse-micelle formation in the partitioning of trivalent f-element cations by biphasic systems containing a tetraalkyldiglycolamide. Langmuir 2007, 23, 47654774,  DOI: 10.1021/la0631926
  25. 25
    Guo, F. Q.; Li, H. F.; Zhang, Z. F.; Meng, S. L.; Li, D. Q. Reversed micelle formation in a model liquid-liquid extraction system. J. Colloid Interface Sci. 2008, 322, 605610,  DOI: 10.1016/j.jcis.2008.03.011
  26. 26
    Chiarizia, R.; Briand, A.; Jensen, M. P.; Thiyagarajan, P. SANS study of reverse micelles formed upon the extraction of inorganic acids by TBP in n-octane. Solvent Extr. Ion Exch. 2008, 26, 333359,  DOI: 10.1080/07366290802182394
  27. 27
    Ganguly, R.; Sharma, J. N.; Choudhury, N. TODGA based w/o microemulsion in dodecane: An insight into the micellar aggregation characteristics by dynamic light scattering and viscometry. J. Colloid Interface Sci. 2011, 355, 458463,  DOI: 10.1016/j.jcis.2010.12.039
  28. 28
    Ellis, R. J.; Meridiano, Y.; Chiarizia, R.; Berthon, L.; Muller, J.; Couston, L.; Antonio, M. R. Periodic behavior of lanthanide coordination within reverse micelles. Chem. - Eur. J. 2013, 19, 26632675,  DOI: 10.1002/chem.201202880
  29. 29
    Guilbaud, P.; Zemb, T. Depletion of water-in-oil aggregates from poor solvents: Transition from weak aggregates towards reverse micelles. Curr. Opin. Colloid Interface Sci. 2015, 20, 7177,  DOI: 10.1016/j.cocis.2014.11.011
  30. 30
    Bley, M.; Siboulet, B.; Karmakar, A.; Zemb, T.; Dufreche, J. F. A predictive model of reverse micelles solubilizing water for solvent extraction. J. Colloid Interface Sci. 2016, 479, 106114,  DOI: 10.1016/j.jcis.2016.06.044
  31. 31
    Diamant, H.; Andelman, D. Free energy approach to micellization and aggregation: Equilibrium, metastability, and kinetics. Curr. Opin. Colloid Interface Sci. 2016, 22, 9498,  DOI: 10.1016/j.cocis.2016.03.004
  32. 32
    Gao, S.; Sun, T. X.; Chen, Q. D.; Shen, X. H. Characterization of reversed micelles formed in solvent extraction of thorium(IV) by bis(2-ethylhexyl) phosphoric acid. Transforming from rodlike to wormlike morphology. Radiochim. Acta 2016, 104, 457469,  DOI: 10.1515/ract-2015-2538
  33. 33
    Chen, Y. S.; Duvail, M.; Guilbaud, P.; Dufreche, J. F. Stability of reverse micelles in rare-earth separation: a chemical model based on a molecular approach. Phys. Chem. Chem. Phys. 2017, 19, 70947100,  DOI: 10.1039/C6CP07843E
  34. 34
    Tonova, K.; Lazarova, Z. Reversed micelle solvents as tools of enzyme purification and enzyme-catalyzed conversion. Biotechnol. Adv. 2008, 26, 516532,  DOI: 10.1016/j.biotechadv.2008.06.002
  35. 35
    Ballesteros-Gomez, A.; Sicilia, M. D.; Rubio, S. Supramolecular solvents in the extraction of organic compounds. A review. Anal. Chim. Acta 2010, 677, 108130,  DOI: 10.1016/j.aca.2010.07.027
  36. 36
    Poirot, R.; Le Goff, X.; Diat, O.; Bourgeois, D.; Meyer, D. Metal recognition driven by weak interactions: A case study in solvent extraction. ChemPhysChem 2016, 17, 21122117,  DOI: 10.1002/cphc.201600305
  37. 37
    Prevost, S.; Gradzielski, M.; Zemb, T. Self-assembly, phase behaviour and structural behaviour as observed by scattering for classical and non-classical microemulsions. Adv. Colloid Interface Sci. 2017, 247, 374396,  DOI: 10.1016/j.cis.2017.07.022
  38. 38
    Moyer, B. A., Ed. Ion Exchange and Solvent Extraction: A Series of Advances; CRC Press: Boca Raton, FL, 2010; p 19.
  39. 39
    Ellis, R. J. Critical exponents for solvent extraction resolved using SAXS. J. Phys. Chem. B 2014, 118, 315322,  DOI: 10.1021/jp408078v
  40. 40
    Plaue, J.; Gelis, A.; Czerwinski, K.; Thiyagarajan, P.; Chiarizia, R. Small-angle neutron scattering study of plutonium third phase formation in 30% TBP/HNO3/alkane diluent systems. Solvent Extr. Solvent Extr. Ion Exch. 2006, 24, 283298,  DOI: 10.1080/07366290600646970
  41. 41
    Chiarizia, R.; Jensen, M. P.; Rickert, P. G.; Kolarik, Z.; Borkowski, M.; Thiyagarajan, P. Extraction of zirconium nitrate by TBP in n-octane: Influence of cation type on third phase formation according to the ″sticky spheres″ model. Langmuir 2004, 20, 1079810808,  DOI: 10.1021/la0488957
  42. 42
    Ivanov, P.; Mu, J.; Leay, L.; Chang, S. Y.; Sharrad, C. A.; Masters, A. J.; Schroeder, S. L. M. Organic and third phase in HNO3/TBP/n-dodecane system: No reverse micelles. Solvent Extr. Ion Exch. 2017, 35, 251265,  DOI: 10.1080/07366299.2017.1336048
  43. 43
    Servis, M.; Wu, D.; Braley, J. Network analysis and percolation transition in hydrogen bonded clusters: nitric acid and water extracted by tributyl phosphate. Phys. Chem. Chem. Phys. 2017, 19, 1132611339,  DOI: 10.1039/C7CP01845B
  44. 44
    Mu, J.; Motokawa, R.; Akutsu, K.; Nishitsuji, S.; Masters, A. J. A novel microemulsion phase transition: Toward the elucidation of third-phase formation in spent nuclear fuel reprocessing. J. Phys. Chem. B 2018, 122, 14391452,  DOI: 10.1021/acs.jpcb.7b08515
  45. 45
    Baldwin, A. G.; Servis, M. J.; Yang, Y.; Bridges, N. J.; Wu, D. T.; Shafer, J. C. The structure of tributyl phosphate solutions: Nitric acid, uranium(VI), and zirconium(IV). J. Mol. Liq. 2017, 246, 225235,  DOI: 10.1016/j.molliq.2017.09.032
  46. 46
    Borkowski, M.; Chiarizia, R.; Jensen, M. P.; Ferraro, J. R.; Thiyagarajan, P.; Littrell, K. C. SANS study of third phase formation in the Th(IV)-HNO3/TBP-n-octane system. Sep. Sci. Technol. 2003, 38, 33333351,  DOI: 10.1081/SS-120022600
  47. 47
    Chiarizia, R.; Jensen, M. P.; Borkowski, M.; Ferraro, J. R.; Thiyagarajan, P.; Littrell, K. C. Third phase formation revisited: The U(VI), HNO3-TBP, n-dodecane system. Solvent Extr. Ion Exch. 2003, 21, 127,  DOI: 10.1081/SEI-120017545
  48. 48
    Motokawa, R.; Endo, H.; Nagao, M.; Heller, W. T. Neutron polarization analysis for biphasic solvent extraction systems. Solvent Extr. Ion Exch. 2016, 34, 399406,  DOI: 10.1080/07366299.2016.1201980
  49. 49
    Koga, T.; Tanaka, F.; Motokawa, R.; Koizumi, S.; Winnik, F. M. Theoretical modeling of associated structures in aqueous solutions of hydrophobically modified telechelic PNIPAM based on a neutron scattering study. Macromolecules 2008, 41, 94139422,  DOI: 10.1021/ma800957z
  50. 50
    Debye, P. Zerstreuung von röntgenstrahlen. Ann. Phys. (Berlin, Ger.) 1915, 351, 809823,  DOI: 10.1002/andp.19153510606
  51. 51
    Svergun, D.; Barberato, C.; Koch, M. H. J. J. CRYSOL– a program to evaluate X-ray solution scattering of biological macromolecules from atomic coordinates. J. Appl. Crystallogr. 1995, 28, 768773,  DOI: 10.1107/S0021889895007047
  52. 52
    Antonio, M. R.; Demars, T. J.; Audras, M.; Ellis, R. J. Third phase inversion, red oil formation, and multinuclear speciation of tetravalent cerium in the tri-n-butyl phosphate-n-dodecane solvent extraction system. Sep. Sci. Technol. 2018, 53, 18341847,  DOI: 10.1080/01496395.2017.1281303
  53. 53
    Antonio, M. R.; Ellis, R. J.; Estes, S. L.; Bera, M. K. Structural insights into the multinuclear speciation of tetravalent cerium in the tri-n-butyl phosphate-n-dodecane solvent extraction system. Phys. Chem. Chem. Phys. 2017, 19, 2130421316,  DOI: 10.1039/C7CP03350H
  54. 54
    Ankudinov, A. L.; Ravel, B.; Rehr, J. J.; Conradson, S. D. Real-space multiple-scattering calculation and interpretation of X-ray-absorption near-edge structure. Phys. Rev. B: Condens. Matter Mater. Phys. 1998, 58, 75657576,  DOI: 10.1103/PhysRevB.58.7565
  55. 55
    Roe, R.-J. Methods of X-ray and Neutron Scattering in Polymer Science, 1st ed.; Oxford University Press: NY, 2000.
  56. 56
    Okuhara, T.; Mizuno, N.; Misono, M. In Adv. Catal.; Eley, D. D., Haag, W. O., Gates, B., Eds.; 1996; Vol. 41, pp 113252.
  57. 57
    Bartlett, P.; Ottewill, R. H. A neutron scattering study of the structure of a bimodal colloidal crystal. J. Chem. Phys. 1992, 96, 33063318,  DOI: 10.1063/1.461926
  58. 58
    Qiao, B.; Demars, T.; Olvela de la Cruz, M.; Ellis, R. J. How hydrogen bonds affect the growth of reverse micelles around coordinating metal ions. J. Phys. Chem. Lett. 2014, 5, 14401444,  DOI: 10.1021/jz500495p
  59. 59
    Ellis, R. J.; Meridiano, Y.; Muller, J.; Berthon, L.; Guilbaud, P.; Zorz, N.; Antonio, M. R.; Demars, T.; Zemb, T. Complexation-induced supramolecular assembly drives metal-ion extraction. Chem. - Eur. J. 2014, 20, 1279612807,  DOI: 10.1002/chem.201403859
  60. 60
    Qiao, B. F.; Ferru, G.; Olvela de la Cruz, M.; Ellis, R. J. Molecular origins of mesoscale ordering in a metalloamphiphile phase. ACS Cent. Sci. 2015, 1, 493503,  DOI: 10.1021/acscentsci.5b00306
  61. 61
    Sevick, E.; Monson, P.; Ottino, J. Monte-carlo calculations of cluster statistics in continuum models of composite morphology. J. Chem. Phys. 1988, 88, 11981206,  DOI: 10.1063/1.454720
  62. 62
    Servis, M.; Wu, D.; Jenifer, S.; Clark, A. Square supramolecular assemblies of uranyl complexes in organic solvents. Chem. Commun. 2018, 54, 1006410067,  DOI: 10.1039/C8CC05277H
  63. 63
    Hill, T. L. An Introduction to Statistical Thermodynamics; Addison-Wesley: Reading, MA, 1960.
  64. 64
    Higgins, J. S.; Benoît, H. C. Polymers and Neutron Scattering; Oxford University Press: Oxford, 1994.
  65. 65
    Berthon, L.; Martinet, L.; Testard, F.; Madic, C.; Zemb, T. Solvent penetration and sterical stabilization of reverse aggregates based on the DIAMEX process extracting molecules: Consequences for the third phase formation. Solvent Extr. Ion Exch. 2007, 25, 545576,  DOI: 10.1080/07366290701512576
  66. 66
    Zemb, T.; Bauer, C.; Bauduin, P.; Belloni, L.; Dejugnat, C.; Diat, O.; Dubois, V.; Dufreche, J. F.; Dourdain, S.; Duvail, M.; Larpent, C.; Testard, F.; Pellet-Rostaing, S. Recycling metals by controlled transfer of ionic species between complex fluids: en route to ″ienaics″. Colloid Polym. Sci. 2015, 293, 122,  DOI: 10.1007/s00396-014-3447-x
  67. 67
    Abécassis, B.; Testard, F.; Zemb, T.; Berthon, L.; Madic, C. Effect of n-octanol on the structure at the supramolecular scale of concentrated dimethyldioctylhexylethoxymalonamide extractant solutions. Langmuir 2003, 19, 66386644,  DOI: 10.1021/la034088g
  68. 68
    Whittaker, D.; Geist, A.; Modolo, G.; Taylor, R.; Sarsfield, M.; Wilden, A. Applications of diglycolamide based solvent extraction processes in spent nuclear fuel reprocessing, Part 1: TODGA. Solvent Extr. Ion Exch. 2018, 36, 223256,  DOI: 10.1080/07366299.2018.1464269

Cited By

This article is cited by 43 publications.

  1. Daniel Massey, Christopher D. Williams, Junju Mu, Andrew J. Masters, Ryuhei Motokawa, Noboru Aoyagi, Yuki Ueda, Mark R. Antonio. Hierarchical Aggregation in a Complex Fluid─The Role of Isomeric Interconversion. The Journal of Physical Chemistry B 2023, 127 (9) , 2052-2065. https://doi.org/10.1021/acs.jpcb.2c07527
  2. Darren M. Driscoll, Hongjun Liu, Benjamin Reinhart, Ilja Popovs, Vera Bocharova, Santa Jansone-Popova, De-en Jiang, Alexander S. Ivanov. Noncoordinating Secondary Sphere Ion Modulates Supramolecular Clustering of Lanthanides. The Journal of Physical Chemistry Letters 2022, 13 (51) , 12076-12081. https://doi.org/10.1021/acs.jpclett.2c03423
  3. Daniel Massey, Andrew Masters, Jonathan Macdonald-Taylor, David Woodhead, Robin Taylor. Molecular Dynamics Study of the Aggregation Behavior of N,N,N′,N′-Tetraoctyl Diglycolamide. The Journal of Physical Chemistry B 2022, 126 (33) , 6290-6300. https://doi.org/10.1021/acs.jpcb.2c02198
  4. Yuki Ueda, Ayano Eguchi, Kohei Tokunaga, Kei Kikuchi, Tsuyoshi Sugita, Hiroyuki Okamura, Hirochika Naganawa. Urea-Introduced Ionic Liquid for the Effective Extraction of Pt(IV) and Pd(II) Ions. Industrial & Engineering Chemistry Research 2022, 61 (19) , 6640-6649. https://doi.org/10.1021/acs.iecr.2c00304
  5. D. Sheyfer, Michael J. Servis, Qingteng Zhang, J. Lal, T. Loeffler, E. M. Dufresne, A. R. Sandy, S. Narayanan, Subramanian K.R.S. Sankaranarayanan, R. Szczygiel, P. Maj, L. Soderholm, Mark R. Antonio, G. B. Stephenson. Advancing Chemical Separations: Unraveling the Structure and Dynamics of Phase Splitting in Liquid–Liquid Extraction. The Journal of Physical Chemistry B 2022, 126 (12) , 2420-2429. https://doi.org/10.1021/acs.jpcb.1c09996
  6. Mario Špadina, Jean-François Dufrêche, Stephane Pellet-Rostaing, Stjepan Marčelja, Thomas Zemb. Molecular Forces in Liquid–Liquid Extraction. Langmuir 2021, 37 (36) , 10637-10656. https://doi.org/10.1021/acs.langmuir.1c00673
  7. Michael J. Servis, Marek Piechowicz, L. Soderholm. Impact of Water Extraction on Malonamide Aggregation: A Molecular Dynamics and Graph Theoretic Approach. The Journal of Physical Chemistry B 2021, 125 (24) , 6629-6638. https://doi.org/10.1021/acs.jpcb.1c02962
  8. Michael J. Servis, G. B. Stephenson. Mesostructuring in Liquid–Liquid Extraction Organic Phases Originating from Critical Points. The Journal of Physical Chemistry Letters 2021, 12 (24) , 5807-5812. https://doi.org/10.1021/acs.jpclett.1c01429
  9. Srikanth Nayak, Raju R. Kumal, Zhu Liu, Baofu Qiao, Aurora E. Clark, Ahmet Uysal. Origins of Clustering of Metalate–Extractant Complexes in Liquid–Liquid Extraction. ACS Applied Materials & Interfaces 2021, 13 (20) , 24194-24206. https://doi.org/10.1021/acsami.0c23158
  10. Michael J. Servis, Aurora E. Clark. Cluster Identification Using Modularity Optimization to Uncover Chemical Heterogeneity in Complex Solutions. The Journal of Physical Chemistry A 2021, 125 (18) , 3986-3993. https://doi.org/10.1021/acs.jpca.0c11320
  11. Michael J. Servis, Marek Piechowicz, Ilya A. Shkrob, L. Soderholm, Aurora E. Clark. Amphiphile Organization in Organic Solutions: An Alternative Explanation for Small-Angle X-ray Scattering Features in Malonamide/Alkane Mixtures. The Journal of Physical Chemistry B 2020, 124 (47) , 10822-10831. https://doi.org/10.1021/acs.jpcb.0c07080
  12. Mario Špadina, Klemen Bohinc, Thomas Zemb, Jean-François Dufrêche. Synergistic Solvent Extraction Is Driven by Entropy. ACS Nano 2019, 13 (12) , 13745-13758. https://doi.org/10.1021/acsnano.9b07605
  13. Hirokazu Narita, Rebecca M. Nicolson, Ryuhei Motokawa, Fumiyuki Ito, Kazuko Morisaku, Midori Goto, Mikiya Tanaka, William T. Heller, Hideaki Shiwaku, Tsuyoshi Yaita, Ross J. Gordon, Jason B. Love, Peter A. Tasker, Emma R. Schofield, Mark R. Antonio, Carole A. Morrison. Proton Chelating Ligands Drive Improved Chemical Separations for Rhodium. Inorganic Chemistry 2019, 58 (13) , 8720-8734. https://doi.org/10.1021/acs.inorgchem.9b01136
  14. Taishi Kobayashi, Shogo Nakajima, Ryuhei Motokawa, Daiju Matsumura, Takumi Saito, Takayuki Sasaki. Structural Approach to Understanding the Solubility of Metal Hydroxides. Langmuir 2019, 35 (24) , 7995-8006. https://doi.org/10.1021/acs.langmuir.9b01132
  15. Aurora E. Clark. Amphiphile-Based Complex Fluids: The Self-Assembly Ensemble as Protagonist. ACS Central Science 2019, 5 (1) , 10-12. https://doi.org/10.1021/acscentsci.8b00927
  16. Allison A. Peroutka, Shane S. Galley, Jenifer C. Shafer. Elucidating the speciation of extracted lanthanides by diglycolamides. Coordination Chemistry Reviews 2023, 482 , 215071. https://doi.org/10.1016/j.ccr.2023.215071
  17. Ueslei G. Favero, Nicolas Schaeffer, Helena Passos, Kaíque A. M. L. Cruz, Duarte Ananias, Sandrine Dourdain, Maria C. Hespanhol. Solvent extraction in non-ideal eutectic solvents – application towards lanthanide separation. Separation and Purification Technology 2023, 337 , 123592. https://doi.org/10.1016/j.seppur.2023.123592
  18. Cyril Micheau, Yuki Ueda, Kazuhiro Akutsu-Suyama, Damien Bourgeois, Ryuhei Motokawa. Deuterated Malonamide Synthesis for Fundamental Research on Solvent Extraction Systems. Solvent Extraction and Ion Exchange 2023, 41 (2) , 221-240. https://doi.org/10.1080/07366299.2023.2166351
  19. Anna P.S. Crema, Nicolas Schaeffer, Henrique Bastos, Liliana P. Silva, Dinis O. Abranches, Helena Passos, Maria C. Hespanhol, João A.P. Coutinho. New family of Type V eutectic solvents based on 1,10-phenanthroline and their application in metal extraction. Hydrometallurgy 2023, 215 , 105971. https://doi.org/10.1016/j.hydromet.2022.105971
  20. Alexander P. Antonov, Sören Schweers, Artem Ryabov, Philipp Maass. Brownian dynamics simulations of hard rods in external fields and with contact interactions. Physical Review E 2022, 106 (5) https://doi.org/10.1103/PhysRevE.106.054606
  21. Subramee Sarkar, A. Suresh, N. Sivaraman, Vinod K Aswal. Studies on the aggregation behavior of mineral acid and Zr(IV) loaded Tris(2-methylbutyl) phosphate and tri- n -alkyl phosphate systems using small angle neutron scattering. Separation Science and Technology 2022, 57 (6) , 968-978. https://doi.org/10.1080/01496395.2021.1954021
  22. Zijun Lu, Sandrine Dourdain, Jean-François Dufrêche, Bruno Demé, Thomas Zemb, Stéphane Pellet-Rostaing. Effect of alkyl chains configurations of tertiary amines on uranium extraction and phase stability – part II: Curvature free energy controlling the ion transfer. Journal of Molecular Liquids 2022, 349 , 118487. https://doi.org/10.1016/j.molliq.2022.118487
  23. Simon Stemplinger, Magali Duvail, Jean-François Dufrêche. Molecular dynamics simulations of Eu(NO3)3 salt with DMDOHEMA in n-alkanes: Unravelling curvature properties in liquid-liquid extraction. Journal of Molecular Liquids 2022, 348 , 118035. https://doi.org/10.1016/j.molliq.2021.118035
  24. M. Špadina, S. Dourdain, J. Rey, K. Bohinc, S. Pellet-Rostaing, Jean-François Dufrêche, T. Zemb. How acidity rules synergism and antagonism in liquid–liquid extraction by lipophilic extractants—Part II: application of the ienaic modelling. Solvent Extraction and Ion Exchange 2022, 40 (1-2) , 106-139. https://doi.org/10.1080/07366299.2021.1899614
  25. Michael J. Servis, Srikanth Nayak, Soenke Seifert. The pervasive impact of critical fluctuations in liquid–liquid extraction organic phases. The Journal of Chemical Physics 2021, 155 (24) , 244506. https://doi.org/10.1063/5.0074995
  26. Norman Kelly, Thomas Doert, Felix Hennersdorf, Karsten Gloe. Synergistic lanthanide extraction triggered by self-assembly of heterodinuclear Zn(II)/Ln(III) Schiff base/carboxylic acid complexes. Solvent Extraction and Ion Exchange 2021, 39 (5-6) , 545-572. https://doi.org/10.1080/07366299.2021.1876383
  27. Yuki Ueda, Shintaro Morisada, Hidetaka Kawakita, Keisuke Ohto. Selective Extraction of Platinum(IV) from the Simulated Secondary Resources Using Simple Secondary Amide and Urea Extractants. Separations 2021, 8 (9) , 139. https://doi.org/10.3390/separations8090139
  28. Nicolas Schaeffer, Silvia J. R. Vargas, Helena Passos, Paula Brandão, Helena I. S. Nogueira, Lenka Svecova, Papaiconomou, João A. P. Coutinho. A HNO 3 ‐Responsive Aqueous Biphasic System for Metal Separation: Application towards Ce IV Recovery. ChemSusChem 2021, 14 (14) , 3018-3026. https://doi.org/10.1002/cssc.202101149
  29. Margarita Kruteva. Dynamics studied by Quasielastic Neutron Scattering (QENS). Adsorption 2021, 27 (5) , 875-889. https://doi.org/10.1007/s10450-020-00295-4
  30. Silvia J. R. Vargas, Germán Pérez-Sánchez, Nicolas Schaeffer, João A. P. Coutinho. Solvent extraction in extended hydrogen bonded fluids – separation of Pt( iv ) from Pd( ii ) using TOPO-based type V DES. Green Chemistry 2021, 23 (12) , 4540-4550. https://doi.org/10.1039/D1GC00829C
  31. Bassam I. El-Eswed, Mahmoud Sunjuk, Raed Ghuneim, Yahya.S. Al-Degs, Maha Al Rimawi, Yanal Albawarshi. Competitive extraction of Li, Na, K, Mg and Ca ions from acidified aqueous solutions into chloroform layer containing diluted alkyl phosphates. Journal of Colloid and Interface Science 2021, 587 , 229-239. https://doi.org/10.1016/j.jcis.2020.12.029
  32. Takuya Okudaira, Yuki Ueda, Kosuke Hiroi, Ryuhei Motokawa, Yasuhiro Inamura, Shin-ichi Takata, Takayuki Oku, Jun-ichi Suzuki, Shingo Takahashi, Hitoshi Endo, Hiroki Iwase. Polarization analysis for small-angle neutron scattering with a 3 He spin filter at a pulsed neutron source. Journal of Applied Crystallography 2021, 54 (2) , 548-556. https://doi.org/10.1107/S1600576721001643
  33. Hirokazu Narita, Ryo Kasuya, Tomoya Suzuki, Ryuhei Motokawa, Mikiya Tanaka. Precious Metal Separations. 2020, 1-28. https://doi.org/10.1002/9781119951438.eibc2756
  34. Srikanth Nayak, Kaitlin Lovering, Ahmet Uysal. Ion-specific clustering of metal–amphiphile complexes in rare earth separations. Nanoscale 2020, 12 (39) , 20202-20210. https://doi.org/10.1039/D0NR04231E
  35. Koichiro Takao, Yasuhisa Ikeda. Coordination Chemistry of Actinide Nitrates with Cyclic Amide Derivatives for the Development of the Nuclear Fuel Materials Selective Precipitation (NUMAP) Reprocessing Method. European Journal of Inorganic Chemistry 2020, 2020 (36) , 3443-3459. https://doi.org/10.1002/ejic.202000504
  36. D. Sheyfer, Qingteng Zhang, J. Lal, T. Loeffler, E. M. Dufresne, A. R. Sandy, S. Narayanan, S. K. R. S. Sankaranarayanan, R. Szczygiel, P. Maj, L. Soderholm, M. R. Antonio, G. B. Stephenson. Nanoscale Critical Phenomena in a Complex Fluid Studied by X-Ray Photon Correlation Spectroscopy. Physical Review Letters 2020, 125 (12) https://doi.org/10.1103/PhysRevLett.125.125504
  37. Yojiro Oba, Ryuhei Motokawa, Masahiro Hino, Nozomu Adachi, Yoshikazu Todaka, Rintaro Inoue, Masaaki Sugiyama. Nanostructural Characterization of Oleyl Acid Phosphate in Poly-α-olefin Using Small-angle X-ray Scattering. Chemistry Letters 2020, 49 (7) , 823-825. https://doi.org/10.1246/cl.200204
  38. Michael J. Servis, Ernesto Martinez-Baez, Aurora E. Clark. Hierarchical phenomena in multicomponent liquids: simulation methods, analysis, chemistry. Physical Chemistry Chemical Physics 2020, 22 (18) , 9850-9874. https://doi.org/10.1039/D0CP00164C
  39. Damien Bourgeois, Asmae El Maangar, Sandrine Dourdain. Importance of weak interactions in the formulation of organic phases for efficient liquid/liquid extraction of metals. Current Opinion in Colloid & Interface Science 2020, 46 , 36-51. https://doi.org/10.1016/j.cocis.2020.03.004
  40. Mario Špadina, Klemen Bohinc. Multiscale modeling of solvent extraction and the choice of reference state: Mesoscopic modeling as a bridge between nanoscale and chemical engineering. Current Opinion in Colloid & Interface Science 2020, 46 , 94-113. https://doi.org/10.1016/j.cocis.2020.03.011
  41. Asmae El Maangar, Johannes Theisen, Christophe Penisson, Thomas Zemb, Jean-Christophe P. Gabriel. A microfluidic study of synergic liquid–liquid extraction of rare earth elements. Physical Chemistry Chemical Physics 2020, 22 (10) , 5449-5462. https://doi.org/10.1039/C9CP06569E
  42. Yuki Ueda, Kei Kikuchi, Tsuyoshi Sugita, Ryuhei Motokawa. Extraction Performance of a Fluorous Phosphate for Zr(IV) from HNO 3 Solution: Comparison with Tri- n -Butyl Phosphate. Solvent Extraction and Ion Exchange 2019, 37 (5) , 347-359. https://doi.org/10.1080/07366299.2019.1638015
  43. J. Durain, D. Bourgeois, M. Bertrand, D. Meyer. Comprehensive Studies on Third Phase Formation: Application to U (VI) /Th (IV) Mixtures Extracted by TBP in N -dodecane. Solvent Extraction and Ion Exchange 2019, 37 (5) , 328-346. https://doi.org/10.1080/07366299.2019.1656853
  • Abstract

    Figure 1

    Figure 1. EXAFS spectra for the extracted Zr coordination complexes, (a) k3-weighted Zr K-edge EXAFS, k3χ(k) (open black circles), and (b) corresponding Fourier transform, |FT[k3χ(k)]| (open black circles), and the imaginary part of FT[k3χ(k)], Im{FT[k3χ(k)]} (filled blue circles), obtained for the organic phases of sample nos. 2–5. The solid black curves in part a, the red curves in part b, and the green curves in part b are the simulated k3χ(k), |FT[k3χ(k)]|, and Im{FT[k3χ(k)]} responses, respectively. The thick arrows highlight the P1, P2, P3, and P4 peaks, which originate from the scattering paths of Zr–OTBP and Zr–ONO3, Zr–NNO3, Zr–PTBP, and Zr–N–Omultiple, respectively (Table 2).

    Figure 2

    Figure 2. Schematic diagrams of hierarchical aggregate model of zirconium superclusters. (a) Geometry of the optimized coordination structure of extracted Zr(NO3)4(TBP)2 in the organic phase, determined by DFT calculation. Green, Zr; yellow, P; red, O; blue, N; black, C; and light pink, H. (b) Primary cluster in which the Zr(NO3)4(TBP)2 complexes (red spheres) distribute with radius RS around the central complex, (c) primary clusters assemble into a large aggregate (supercluster), where the primary clusters with radius RS surround the central cluster (light blue sphere) with radius RL. A set of the number of the primary clusters, M = 25, and the number of the complexes, N = 7, corresponds to the characteristic parameters of sample no. 5 from SANS data analysis.

    Figure 3

    Figure 3. Double-logarithmic plots of the SANS profiles, (a) Iobs(q) and (b) Isub(q), with error bars, as a function of [Zr(NO3)4(TBP)2]org,eq. [Zr(NO3)4(TBP)2]org,eq gradually increases with sample no. from 1 to 5. Dashed lines in part b are the form factors of Zr(NO3)4(TBP)2 in the organic phase, determined on the basis of the Debye scattering formula for randomly orientated Zr(NO3)4(TBP)2 using eqs S3, S6, and S7. Solid lines in part b are the best-fit theoretical SANS profiles obtained by using eqs S10–S13 together with the characteristic parameters listed in Table 3.

    Figure 4

    Figure 4. (a) Zr–Zr radial distribution function, g(rZr–Zr) (solid line), and corresponding coordination number of Zr with the other Zr, CN(rZr–Zr) (dashed line). (b) The primary cluster probability distribution for system no. 5 as determined by the MD. The ordinate gives the probability of finding a primary cluster with a given number of Zr atoms. This probability is the number of primary clusters of a given aggregation number divided by the total number of the primary cluster.

    Figure 5

    Figure 5. (a) Snapshot of two, neighboring primary clusters from our MD simulations and (b) magnified snapshot showing the hydrogen-bonding network within the primary clusters. Green, Zr; yellow, P; red, O; blue, N; black, C; light pink, H; and light blue dashed line, hydrogen bond.

    Figure 6

    Figure 6. (a) Number of hydrogen bonds per Zr complex as a function of aggregation number. (b) Radial distribution function for octane carbon atoms around a central Zr atom as a function of cluster aggregation number of Zr complexes per primary cluster. (c) Radial distribution function for octane carbon atoms around a central N atom in a nitrate ligand, shown as a function of cluster aggregation number. All results in parts a–c are obtained for system no. 5.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 68 other publications.

    1. 1
      Gompper, G.; Schick, M. In Phase Transitions and Critical Phenomena; Domb, C., Lebowitz, J., Eds.; Academic Press: London, 1994; Vol. 16, pp 1176.
    2. 2
      Schwuger, M. J.; Stickdorn, K.; Schomacker, R. Microemulsions in technical processes. Chem. Rev. 1995, 95, 849864,  DOI: 10.1021/cr00036a003
    3. 3
      Winsor, P. A. Solvent Properties of Amphiphilic Compounds; Butterworths: London, 1954.
    4. 4
      Lindman, B.; Wennerström, H. Miceles. Amphiphile aggregation in aqueous solution. Top. Curr. Chem. 1980, 87, 183,  DOI: 10.1007/BFb0048488
    5. 5
      Eicke, H.-F. Surfactants in nonpolar solvents. Aggregation and micellization. Top. Curr. Chem. 1980, 87, 85145,  DOI: 10.1007/BFb0048489
    6. 6
      Fischer, S.; Exner, A.; Zielske, K.; Perlich, J.; Deloudi, S.; Steurer, W.; Lindner, P.; Forster, S. Colloidal quasicrystals with 12-fold and 18-fold diffraction symmetry. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 18101814,  DOI: 10.1073/pnas.1008695108
    7. 7
      Zemb, T. Flexibility, persistence length and bicontinuous microstructures in microemulsions. C. R. Chim. 2009, 12, 218224,  DOI: 10.1016/j.crci.2008.10.008
    8. 8
      Zemb, T.; Holmberg, K.; Kunz, W. Special topic: Weak self assembly. Curr. Opin. Colloid Interface Sci. 2016, 22, A1A3,  DOI: 10.1016/j.cocis.2016.04.001
    9. 9
      Zemb, T.; Kunz, W. Weak aggregation: State of the art, expectations and open questions. Curr. Opin. Colloid Interface Sci. 2016, 22, 113119,  DOI: 10.1016/j.cocis.2016.04.002
    10. 10
      Pini, D.; Parola, A. Pattern formation and self-assembly driven by competing interactions. Soft Matter 2017, 13, 92599272,  DOI: 10.1039/C7SM02125A
    11. 11
      Foffi, G.; De Michele, C.; Sciortino, F.; Tartaglia, P. Scaling of dynamics with the range of interaction in short-range attractive colloids. Phys. Rev. Lett. 2005, 94, 078301,  DOI: 10.1103/PhysRevLett.94.078301
    12. 12
      Zaccarelli, E. Colloidal gels: equilibrium and non-equilibrium routes. J. Phys.: Condens. Matter 2007, 19, 323101,  DOI: 10.1088/0953-8984/19/32/323101
    13. 13
      Bera, M. K.; Qiao, B. F.; Seifert, S.; Burton-Pye, B. P.; Olvela de la Cruz, M.; Antonio, M. R. Aggregation of heteropolyanions in aqueous solutions exhibiting short-range attractions and long-range repulsions. J. Phys. Chem. C 2016, 120, 13171327,  DOI: 10.1021/acs.jpcc.5b10609
    14. 14
      Poon, W. C. K.; Haw, M. D. Mesoscopic structure formation in colloidal aggregation and gelation. Adv. Colloid Interface Sci. 1997, 73, 71126,  DOI: 10.1016/S0001-8686(97)90003-8
    15. 15
      Lu, P. J.; Zaccarelli, E.; Ciulla, F.; Schofield, A. B.; Sciortino, F.; Weitz, D. A. Gelation of particles with short-range attraction. Nature 2008, 453, 499505,  DOI: 10.1038/nature06931
    16. 16
      Kim, S. A.; Jeong, K. J.; Yethiraj, A.; Mahanthappa, M. K. Low-symmetry sphere packings of simple surfactant micelles induced by ionic sphericity. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 40724077,  DOI: 10.1073/pnas.1701608114
    17. 17
      Bauer, C.; Bauduin, P.; Dufreche, J. F.; Zemb, T.; Diat, O. Liquid/liquid metal extraction: Phase diagram topology resulting from molecular interactions between extractant, ion, oil and water. Eur. Phys. J.: Spec. Top. 2012, 213, 225241,  DOI: 10.1140/epjst/e2012-01673-4
    18. 18
      Erlinger, C.; Gazeau, D.; Zemb, T.; Madic, C.; Lefrancois, L.; Hebrant, M.; Tondre, C. Effect of nitric acid extraction on phase behavior, microstructure and interactions between primary aggregates in the system dimethyldibutyltetradecylmalonamide (DMDBTDMA)/n-dodecane/water: A phase analysis and small angle X-ray scattering (SAXS) characterisation study. Solvent Extr. Ion Exch. 1998, 16, 707738,  DOI: 10.1080/07366299808934549
    19. 19
      Testard, F.; Zemb, T.; Bauduin, P.; Berthon, L. In Ion Exchange and Solvent Extraction: A Series of Advances; Moyer, B. A., Ed.; CRC Press: Boca Raton, FL, 2010; Vol. 19, pp 381428.
    20. 20
      Baxter, R. J. Percus-Yevick equation for hard spheres with surface adhesion. J. Chem. Phys. 1968, 49, 27702774,  DOI: 10.1063/1.1670482
    21. 21
      Schulz, W. W.; Navratil, J. D. Science and Technology of Tributyl Phosphate; CRC Press: Boca Raton, FL, 1984.
    22. 22
      Osseo-Asare, K. Aggregation, reversed micelles, and microemulsions in liquid-liquid extraction: the tri-n-butyl phosphate-diluent-water-electrolyte system. Adv. Colloid Interface Sci. 1991, 37, 123173,  DOI: 10.1016/0001-8686(91)80041-H
    23. 23
      Diss, R.; Wipff, G. Lanthanide cation extraction by malonamide ligands: from liquid-liquid interfaces to microemulsions. A molecular dynamics study. Phys. Chem. Chem. Phys. 2005, 7, 264272,  DOI: 10.1039/B410137E
    24. 24
      Jensen, M. P.; Yaita, T.; Chiarizia, R. Reverse-micelle formation in the partitioning of trivalent f-element cations by biphasic systems containing a tetraalkyldiglycolamide. Langmuir 2007, 23, 47654774,  DOI: 10.1021/la0631926
    25. 25
      Guo, F. Q.; Li, H. F.; Zhang, Z. F.; Meng, S. L.; Li, D. Q. Reversed micelle formation in a model liquid-liquid extraction system. J. Colloid Interface Sci. 2008, 322, 605610,  DOI: 10.1016/j.jcis.2008.03.011
    26. 26
      Chiarizia, R.; Briand, A.; Jensen, M. P.; Thiyagarajan, P. SANS study of reverse micelles formed upon the extraction of inorganic acids by TBP in n-octane. Solvent Extr. Ion Exch. 2008, 26, 333359,  DOI: 10.1080/07366290802182394
    27. 27
      Ganguly, R.; Sharma, J. N.; Choudhury, N. TODGA based w/o microemulsion in dodecane: An insight into the micellar aggregation characteristics by dynamic light scattering and viscometry. J. Colloid Interface Sci. 2011, 355, 458463,  DOI: 10.1016/j.jcis.2010.12.039
    28. 28
      Ellis, R. J.; Meridiano, Y.; Chiarizia, R.; Berthon, L.; Muller, J.; Couston, L.; Antonio, M. R. Periodic behavior of lanthanide coordination within reverse micelles. Chem. - Eur. J. 2013, 19, 26632675,  DOI: 10.1002/chem.201202880
    29. 29
      Guilbaud, P.; Zemb, T. Depletion of water-in-oil aggregates from poor solvents: Transition from weak aggregates towards reverse micelles. Curr. Opin. Colloid Interface Sci. 2015, 20, 7177,  DOI: 10.1016/j.cocis.2014.11.011
    30. 30
      Bley, M.; Siboulet, B.; Karmakar, A.; Zemb, T.; Dufreche, J. F. A predictive model of reverse micelles solubilizing water for solvent extraction. J. Colloid Interface Sci. 2016, 479, 106114,  DOI: 10.1016/j.jcis.2016.06.044
    31. 31
      Diamant, H.; Andelman, D. Free energy approach to micellization and aggregation: Equilibrium, metastability, and kinetics. Curr. Opin. Colloid Interface Sci. 2016, 22, 9498,  DOI: 10.1016/j.cocis.2016.03.004
    32. 32
      Gao, S.; Sun, T. X.; Chen, Q. D.; Shen, X. H. Characterization of reversed micelles formed in solvent extraction of thorium(IV) by bis(2-ethylhexyl) phosphoric acid. Transforming from rodlike to wormlike morphology. Radiochim. Acta 2016, 104, 457469,  DOI: 10.1515/ract-2015-2538
    33. 33
      Chen, Y. S.; Duvail, M.; Guilbaud, P.; Dufreche, J. F. Stability of reverse micelles in rare-earth separation: a chemical model based on a molecular approach. Phys. Chem. Chem. Phys. 2017, 19, 70947100,  DOI: 10.1039/C6CP07843E
    34. 34
      Tonova, K.; Lazarova, Z. Reversed micelle solvents as tools of enzyme purification and enzyme-catalyzed conversion. Biotechnol. Adv. 2008, 26, 516532,  DOI: 10.1016/j.biotechadv.2008.06.002
    35. 35
      Ballesteros-Gomez, A.; Sicilia, M. D.; Rubio, S. Supramolecular solvents in the extraction of organic compounds. A review. Anal. Chim. Acta 2010, 677, 108130,  DOI: 10.1016/j.aca.2010.07.027
    36. 36
      Poirot, R.; Le Goff, X.; Diat, O.; Bourgeois, D.; Meyer, D. Metal recognition driven by weak interactions: A case study in solvent extraction. ChemPhysChem 2016, 17, 21122117,  DOI: 10.1002/cphc.201600305
    37. 37
      Prevost, S.; Gradzielski, M.; Zemb, T. Self-assembly, phase behaviour and structural behaviour as observed by scattering for classical and non-classical microemulsions. Adv. Colloid Interface Sci. 2017, 247, 374396,  DOI: 10.1016/j.cis.2017.07.022
    38. 38
      Moyer, B. A., Ed. Ion Exchange and Solvent Extraction: A Series of Advances; CRC Press: Boca Raton, FL, 2010; p 19.
    39. 39
      Ellis, R. J. Critical exponents for solvent extraction resolved using SAXS. J. Phys. Chem. B 2014, 118, 315322,  DOI: 10.1021/jp408078v
    40. 40
      Plaue, J.; Gelis, A.; Czerwinski, K.; Thiyagarajan, P.; Chiarizia, R. Small-angle neutron scattering study of plutonium third phase formation in 30% TBP/HNO3/alkane diluent systems. Solvent Extr. Solvent Extr. Ion Exch. 2006, 24, 283298,  DOI: 10.1080/07366290600646970
    41. 41
      Chiarizia, R.; Jensen, M. P.; Rickert, P. G.; Kolarik, Z.; Borkowski, M.; Thiyagarajan, P. Extraction of zirconium nitrate by TBP in n-octane: Influence of cation type on third phase formation according to the ″sticky spheres″ model. Langmuir 2004, 20, 1079810808,  DOI: 10.1021/la0488957
    42. 42
      Ivanov, P.; Mu, J.; Leay, L.; Chang, S. Y.; Sharrad, C. A.; Masters, A. J.; Schroeder, S. L. M. Organic and third phase in HNO3/TBP/n-dodecane system: No reverse micelles. Solvent Extr. Ion Exch. 2017, 35, 251265,  DOI: 10.1080/07366299.2017.1336048
    43. 43
      Servis, M.; Wu, D.; Braley, J. Network analysis and percolation transition in hydrogen bonded clusters: nitric acid and water extracted by tributyl phosphate. Phys. Chem. Chem. Phys. 2017, 19, 1132611339,  DOI: 10.1039/C7CP01845B
    44. 44
      Mu, J.; Motokawa, R.; Akutsu, K.; Nishitsuji, S.; Masters, A. J. A novel microemulsion phase transition: Toward the elucidation of third-phase formation in spent nuclear fuel reprocessing. J. Phys. Chem. B 2018, 122, 14391452,  DOI: 10.1021/acs.jpcb.7b08515
    45. 45
      Baldwin, A. G.; Servis, M. J.; Yang, Y.; Bridges, N. J.; Wu, D. T.; Shafer, J. C. The structure of tributyl phosphate solutions: Nitric acid, uranium(VI), and zirconium(IV). J. Mol. Liq. 2017, 246, 225235,  DOI: 10.1016/j.molliq.2017.09.032
    46. 46
      Borkowski, M.; Chiarizia, R.; Jensen, M. P.; Ferraro, J. R.; Thiyagarajan, P.; Littrell, K. C. SANS study of third phase formation in the Th(IV)-HNO3/TBP-n-octane system. Sep. Sci. Technol. 2003, 38, 33333351,  DOI: 10.1081/SS-120022600
    47. 47
      Chiarizia, R.; Jensen, M. P.; Borkowski, M.; Ferraro, J. R.; Thiyagarajan, P.; Littrell, K. C. Third phase formation revisited: The U(VI), HNO3-TBP, n-dodecane system. Solvent Extr. Ion Exch. 2003, 21, 127,  DOI: 10.1081/SEI-120017545
    48. 48
      Motokawa, R.; Endo, H.; Nagao, M.; Heller, W. T. Neutron polarization analysis for biphasic solvent extraction systems. Solvent Extr. Ion Exch. 2016, 34, 399406,  DOI: 10.1080/07366299.2016.1201980
    49. 49
      Koga, T.; Tanaka, F.; Motokawa, R.; Koizumi, S.; Winnik, F. M. Theoretical modeling of associated structures in aqueous solutions of hydrophobically modified telechelic PNIPAM based on a neutron scattering study. Macromolecules 2008, 41, 94139422,  DOI: 10.1021/ma800957z
    50. 50
      Debye, P. Zerstreuung von röntgenstrahlen. Ann. Phys. (Berlin, Ger.) 1915, 351, 809823,  DOI: 10.1002/andp.19153510606
    51. 51
      Svergun, D.; Barberato, C.; Koch, M. H. J. J. CRYSOL– a program to evaluate X-ray solution scattering of biological macromolecules from atomic coordinates. J. Appl. Crystallogr. 1995, 28, 768773,  DOI: 10.1107/S0021889895007047
    52. 52
      Antonio, M. R.; Demars, T. J.; Audras, M.; Ellis, R. J. Third phase inversion, red oil formation, and multinuclear speciation of tetravalent cerium in the tri-n-butyl phosphate-n-dodecane solvent extraction system. Sep. Sci. Technol. 2018, 53, 18341847,  DOI: 10.1080/01496395.2017.1281303
    53. 53
      Antonio, M. R.; Ellis, R. J.; Estes, S. L.; Bera, M. K. Structural insights into the multinuclear speciation of tetravalent cerium in the tri-n-butyl phosphate-n-dodecane solvent extraction system. Phys. Chem. Chem. Phys. 2017, 19, 2130421316,  DOI: 10.1039/C7CP03350H
    54. 54
      Ankudinov, A. L.; Ravel, B.; Rehr, J. J.; Conradson, S. D. Real-space multiple-scattering calculation and interpretation of X-ray-absorption near-edge structure. Phys. Rev. B: Condens. Matter Mater. Phys. 1998, 58, 75657576,  DOI: 10.1103/PhysRevB.58.7565
    55. 55
      Roe, R.-J. Methods of X-ray and Neutron Scattering in Polymer Science, 1st ed.; Oxford University Press: NY, 2000.
    56. 56
      Okuhara, T.; Mizuno, N.; Misono, M. In Adv. Catal.; Eley, D. D., Haag, W. O., Gates, B., Eds.; 1996; Vol. 41, pp 113252.
    57. 57
      Bartlett, P.; Ottewill, R. H. A neutron scattering study of the structure of a bimodal colloidal crystal. J. Chem. Phys. 1992, 96, 33063318,  DOI: 10.1063/1.461926
    58. 58
      Qiao, B.; Demars, T.; Olvela de la Cruz, M.; Ellis, R. J. How hydrogen bonds affect the growth of reverse micelles around coordinating metal ions. J. Phys. Chem. Lett. 2014, 5, 14401444,  DOI: 10.1021/jz500495p
    59. 59
      Ellis, R. J.; Meridiano, Y.; Muller, J.; Berthon, L.; Guilbaud, P.; Zorz, N.; Antonio, M. R.; Demars, T.; Zemb, T. Complexation-induced supramolecular assembly drives metal-ion extraction. Chem. - Eur. J. 2014, 20, 1279612807,  DOI: 10.1002/chem.201403859
    60. 60
      Qiao, B. F.; Ferru, G.; Olvela de la Cruz, M.; Ellis, R. J. Molecular origins of mesoscale ordering in a metalloamphiphile phase. ACS Cent. Sci. 2015, 1, 493503,  DOI: 10.1021/acscentsci.5b00306
    61. 61
      Sevick, E.; Monson, P.; Ottino, J. Monte-carlo calculations of cluster statistics in continuum models of composite morphology. J. Chem. Phys. 1988, 88, 11981206,  DOI: 10.1063/1.454720
    62. 62
      Servis, M.; Wu, D.; Jenifer, S.; Clark, A. Square supramolecular assemblies of uranyl complexes in organic solvents. Chem. Commun. 2018, 54, 1006410067,  DOI: 10.1039/C8CC05277H
    63. 63
      Hill, T. L. An Introduction to Statistical Thermodynamics; Addison-Wesley: Reading, MA, 1960.
    64. 64
      Higgins, J. S.; Benoît, H. C. Polymers and Neutron Scattering; Oxford University Press: Oxford, 1994.
    65. 65
      Berthon, L.; Martinet, L.; Testard, F.; Madic, C.; Zemb, T. Solvent penetration and sterical stabilization of reverse aggregates based on the DIAMEX process extracting molecules: Consequences for the third phase formation. Solvent Extr. Ion Exch. 2007, 25, 545576,  DOI: 10.1080/07366290701512576
    66. 66
      Zemb, T.; Bauer, C.; Bauduin, P.; Belloni, L.; Dejugnat, C.; Diat, O.; Dubois, V.; Dufreche, J. F.; Dourdain, S.; Duvail, M.; Larpent, C.; Testard, F.; Pellet-Rostaing, S. Recycling metals by controlled transfer of ionic species between complex fluids: en route to ″ienaics″. Colloid Polym. Sci. 2015, 293, 122,  DOI: 10.1007/s00396-014-3447-x
    67. 67
      Abécassis, B.; Testard, F.; Zemb, T.; Berthon, L.; Madic, C. Effect of n-octanol on the structure at the supramolecular scale of concentrated dimethyldioctylhexylethoxymalonamide extractant solutions. Langmuir 2003, 19, 66386644,  DOI: 10.1021/la034088g
    68. 68
      Whittaker, D.; Geist, A.; Modolo, G.; Taylor, R.; Sarsfield, M.; Wilden, A. Applications of diglycolamide based solvent extraction processes in spent nuclear fuel reprocessing, Part 1: TODGA. Solvent Extr. Ion Exch. 2018, 36, 223256,  DOI: 10.1080/07366299.2018.1464269
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acscentsci.8b00669.

    • Method section and additional data and figures including cluster probabilities, configurations, radial distribution function, chemical structure, EXAFS data, Guinier plots, theoretical modeling of the hierarchical aggregates for SANS data analysis, simulated SANS profiles, and coordination structure of zirconium nitrate complex in initial aqueous phase (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect

This website uses cookies to improve your user experience. By continuing to use the site, you are accepting our use of cookies. Read the ACS privacy policy.

CONTINUE