Role of Ligand Straining in Complexation of Eu3+ ... - ACS Publications

Oct 12, 2018 - Role of Ligand Straining in Complexation of Eu3+–Am3+ Ions by TPEN and PPDEN, Scalar Relativistic DFT Exploration in Conjunction with...
0 downloads 0 Views 4MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2018, 3, 13104−13116

http://pubs.acs.org/journal/acsodf

Role of Ligand Straining in Complexation of Eu3+−Am3+ Ions by TPEN and PPDEN, Scalar Relativistic DFT Exploration in Conjunction with COSMO-RS Sk. Musharaf Ali* Chemical Engineering Division, Bhabha Atomic Research Centre, HBNI, Mumbai 40085, India

ACS Omega 2018.3:13104-13116. Downloaded from pubs.acs.org by 5.62.157.132 on 10/16/18. For personal use only.

S Supporting Information *

ABSTRACT: To search for new ligands suitable for the separation of minor actinides (MA) from lanthanides (Ln) in nuclear waste reprocessing, theoretical (density functional theory) studies were carried out on the complexation (structures, bonding, and thermodynamics) of La3+, Sm3+, Eu3+, and Am3+ complexes with moderately soft donor ligands TPEN [N,N,N′,N′-tetrakis(2-pyridylmethyl)ethylenediamine] and PPDEN [N,N,N′,N″,N″-pentakis(2-pyridylmethyl) diethylenetriamine] in aqueous and nitrobenzene solutions. B3LYP level of theory was used in conjunction with the conductor-like screening model for real systems (COSMO-RS). The metal ions in [M(NO3)2(TPEN)]NO3 and [M(NO3)(PPDEN)](NO3)2 complexes were deca-coordinated with both TPEN and PPDEN. The enthalpy of the complexation with TPEN in an aqueous solution was found to be negative, indicating the exothermic nature of the reaction as observed in the experiments. The calculated values of free energy of complexation follow the experimental trend: Am3+ > Sm3+ > La3+. Furthermore, the calculated free energy with PPDEN is reduced compared to that with TPEN, which may be attributed to the ligand straining during complex formation, which is also reflected in greater residual charges on both the Eu3+ and Am3+ central ions in the complexes of octadentate PPDEN compared to hexadentate TPEN. The experimental complexation selectivity of Am3+ over Eu3+ with TPEN is established by employing COSMO-RS. Furthermore, TPEN is Am3+-selective, whereas PPDEN is Eu3+-selective, which could be exploited for the efficient separation of MA from Ln.

1. INTRODUCTION There is a continuing endeavor in the search of an appropriate ligand molecule for the separation of lanthanides (Ln) from actinides in nuclear waste reprocessing using liquid−liquid extraction.1,2 To reduce the radiotoxicity, it is mandatory to perform the partitioning followed by transmutation of longlived actinides, which requires the separation of actinides from Ln.3 The separation of the minor actinides (MA) from Ln is a major challenge in nuclear waste reprocessing because of the chemical similarities between Ln and actinides, which makes the separation very exigent. Recently, heterocyclic ligands with moderately soft N atoms, which show higher affinity for the actinides, are tested and used for the separation of actinides and Ln.4−6 Among those synthesized and tested are: tridentate 2,6-bis(1,2,4-triazine-3yl)pyridine (BTP), quadridentate 6,6bis(1,2,4-triazin-3-yl)-2,2-bipyridine ligands (BTBPs), 2,9-bis(1,2,4-triazin-3-yl)-1,10-phenanthraline (BTPhen), and hexadentate TPEN [N,N,N,N-tetrakis(2-pyridylmethyl)ethylenediamine].7−11 The process of MA recovery comprised two steps: the concurrent removal of Ln and MA from highlevel liquid waste, followed by the separation of MA from Ln. Though the use of extractants like carbamoylmethylphosphine oxide or tetraoctyl diglycolamide is ascertained in the first step, an appropriate ligand for the second step is yet to be © 2018 American Chemical Society

established, and there is an ongoing search to pick up the suitable one. The exploration for suitable ligands with soft donor atoms for the separation of MA from Ln is a contemporary area of research, as ligands with soft donor atoms, such as nitrogen and sulfur, have the preferable affinity toward MA. Researchers in the European countries are working with the nitrogen donor-based BTP, whereas researchers in China and USA are involved in the research with sulfur-based Cyanex 301. However, both these extractants have a serious disadvantagetheir low chemical stability. Recently, attention has been focused on a hexadentate ligand called TPEN7−11 because of its high chemical stability and nongeneration of secondary wastes by incineration, as this ligand consists of C, H, O, and N atoms, merely. The maximum separation factor SFAm−Eu was shown to be 200 [with 1 mM TPEN at pH 4 in nitrobenzene (NB)−water], which is significantly higher than those reported previously from the DTPA processes12 and thus shows promise in nuclear waste remediation. However, there are some demerits also−− TPEN has a considerable aqueous phase solubility, and its Received: May 10, 2018 Accepted: September 28, 2018 Published: October 12, 2018 13104

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

extended X-ray absorption fine structure (EXAFS) experiments.30,31 The minimum-energy structures using both the COSMO and COSMO-RS approaches of the hydrated clusters of La3+, Sm3+, Eu3+, and Am3+ are displayed in Figure S1, and the calculated structural parameters are presented in Table S1. The average M−O distances using both COSMO-RS and COSMO were found to be in fair agreement with the experimentally determined values using EXAFS spectroscopy30,31 as well as the earlier computed results32−34 (see Table S1). Overall, the calculated average bond distance was found to follow the increasing trend: Eu3+ < Sm3+ < Am3+ < La3+ as per the experimental findings and thus confirms the acceptance of the present computation method, which is subsequently used for the study of the complexes of TPEN and PPDEN in the next section. It is noted that the calculated M−O distances using COSMO-RS are always longer than that using COSMO.35,36 2.1.2. Equilibrium Structure of TPEN and PPDEN. The optimized geometries of the free TPEN and PPDEN ligands using the COSMO-RS approach are displayed in Figure 1a,b

extraction ability is greatly reduced at an acidic pH, that is, pH < 2. Nevertheless, there is a scope of improvement by tailoring the molecule to address the solubility and pH tolerance. Parallel to the experimental studies, plentiful theoretical studies13−20 have been performed to understand the actinide− Ln selectivity through the molecular-level insights by studying the structure, bonding, interaction, and thermodynamics of the metal ligand in the gas as well as in the solution phase. Though theoretical studies have been reported for BTP,21 BTBP,22−25 BTPhen,26 and others,16,17,27,28 the theoretical understanding of complexation of TPEN with An/Ln has just begun.29 TPEN is a hexanitrogen donor, which is basically a 2-pyridylmethyl analogue of ethylenediaminetetraacetic acid, H4EDTA. Recently,29 the complexation of TPEN with An/Ln has been reported using a gas-phase optimized geometry, followed by a single-point energy calculation employing the conductor-like screening model (COSMO), a solvation model, but investigation on TPEN and PPDEN [N,N,N,N,N-pentakis(2pyridylmethyl)diethylenetriamine] (Chart 1) for the complexChart 1. Structure of the Studied Ligands

Figure 1. Optimized structures of free (a) TPEN and (b) PPDEN ligands using the COSMO-RS approach.

ation selectivity of MA and Ln using the COSMO-real systems (COSMO-RS) model has not been reported. The tolerance of TPEN at high acidity by introducing a suitable group to the pyridine rings was discussed.29 Therefore, the objective of this study is to investigate the complexation of the hexadentate TPEN with Ln and actinides using density functional theory (DFT) with a higher level of solvation model like COSMO-RS and comparison with COSMO. Additionally, a higher derivative of TPEN with eight N donors, namely PPDEN, has been newly designed to impart hydrophobicity and has been investigated for its complexation ability and selectivity. Many similar theoretical studies were performed so far using the gas-phase optimized geometry, followed by the single-point energy calculation in solution.14−28 Also, most of the reported results used the gas-phase entropy,14−28 though it is known that the entropy in solution differs considerably from that in the gas phase. In view of this, all the calculations were performed with the structures optimized in the solvent, and the thermodynamic parameters were also calculated for the solvent phase, using the novel COSMO-RS model.

whereas the COSMO-optimized structures are shown in Figure S2. From the calculated structural parameters, it is seen that the optimized structures using both COSMO and COSMO-RS are nearly identical. In TPEN, there are four donor “N” atoms from the four pyridine units and two donor “N” atoms from the ethylene diamine linker, thus making it act as a hexadentate ligand, whereas, in case of PPDEN, there are five donor “N” atoms from the five pyridine units and three donor “N” atoms from the diethylene triamine linker, making it an octadentate. From Figure 1, it is seen that all the donor atoms are not in the same plane or in the same direction. The calculated values of the C−C (1.53 Å) and C−N (1.46 Å) bond lengths of the amine chain of PPDEN were seen to be very close to the experimental results of 1.53 and 1.468 Å, respectively.37 2.1.3. Equilibrium Structures of the Complexes of Metal Ions with TPEN in Water. The optimized structures of the La3+−TPEN complex, [La(NO3)2(TPEN)]NO3, using both COSMO and COSMO-RS are displayed in Figure 2, whereas those of the complexes of Sm3+, Eu3+, and Am3+ are displayed in Figures S3−S5. The calculated bond parameters are listed in Table S2. The coordination about the central metal is that of a bicapped square antiprism (coordination number 10) with six nitrogen atoms from one TPEN and two pairs of oxygen atoms from the bidentate NO3− ions. One uncoordinated NO3− anion remains in the outer sphere to satisfy charge neutrality. From Figure 2, it is seen that four donor “N” atoms from the four pyridine units and two donor “N” atoms from ethylene diamine are coordinated to the central metal ion in the first sphere of coordination. In the case of the COSMO-RSoptimized structure, the average La/Sm−Npy distance was found to be 2.770/2.692 Å, whereas the average La/Sm−Namine

2. RESULTS AND DISCUSSION 2.1. Structural Parameters. A reasonable structure of a chemical species is very important for further energetic calculations. Hence, the various chemical species involved in the present study are optimized, and the calculated results are compared with the reported literature and the available experimental results. As most of the chemical reactions take place in solution and metal ions remain strongly hydrated, first, the hydrated metal ions are optimized. 2.1.1. Microhydrated Metal Ions. All the metal ions studied here were taken to be nonahydrated, as reported in the 13105

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

Figure 2. Optimized structure of the complex of La3+ ion with TPEN in the presence of nitrate ions using (a) COSMO and (b) COSMO-RS approaches.

complex in both the COSMO and COSMO-RS structures. The average Eu/Am−Npy was found to be 2.754/2.778 Å, whereas the Eu/Am−Namine length was found to be 2.863/ 2.826 Å. The average Eu/Am−ONO3− was found to be 2.55/ 2.60 Å, much shorter than the Eu/Am−N bond length. For the COSMO-RS structure, the average Eu/Am−Npy was found to be 2.772/2.764 Å, whereas the Eu/Am−Namine bond length was found to be 2.83/2.88 Å. The average Eu/Am−ONO3 bond length was found to be 2.58/2.62 Å. The bond length predicted using COSMO-RS was seen to be little longer than that predicted using COSMO, except for the Eu/Am−Npy length. It is worth mentioning that the calculated bond length of the Eu−PPDEN complex is longer than that of the Eu− TPEN complex, perhaps because of the ligand straining during complexation. Further, the calculated M−O bond length using COSMO-RS is seen to be slightly longer than that using COSMO. The various M−L bond distances for the Am ion were seen to be longer than that of the Eu ion, as observed with TPEN. It is noted that the complex of Eu3+ ion with PPDEN has been reported but could not be compared because of the lack of access to the published data.40 To calculate the aqueous−organic (NB) biphasic metal ion selectivity, free TPEN and its complexes with the Eu3+ and Am3+ ions in the presence of nitrate co-anions were further reoptimized in NB. The optimized structures of the complexes with TPEN using both COSMO and COSMO-RS are displayed in Figures S7a,b and S8a,b. The calculated structural parameters are given in Table S4. There were no significant changes in the optimized structures in NB. From the overall structural analysis, it is observed that the various structural parameters of the hydrated metal ion and metal ion complexes are in fair agreement with the available experimental results and thus confirm the acceptance of the computational protocol used in this study. 2.2. Thermodynamics. 2.2.1. Hydration Free Energy of the Metal Ions. The accurate free energy of hydration, ΔGhyd, for a metal ion is very important to determine the metal ion− ligand complexation ability as well as for the correct prediction of the metal-ion selectivity. The ΔGhyd values for all the metal ions have been calculated using both COSMO and COSMORS, and the values are presented in Table S5. The calculated ΔGhyd of the La3+, Sm3+, and Eu3+ ions with COSMO-RS is in good agreement with the experimental results41,42 compared to COSMO. It is to be noted that though the value of ΔGhyd of La3+ using COSMO-RS is little lower than that reported from CCSD,43 it is quite better than the value calculated with MP2,32,42 BP86,31 and B3LYP.31 The value of ΔGhyd of the Sm3+ ion using COSMO-RS is better than that reported from

distance was found to be 2.765/2.695 Å. The calculated La/ Sm−N distance was found to be very close to the X-ray experimental bond distance of 2.72/2.636 Å.33,38 The average La/Sm−ONO3 bond length was found to be 2.727/2.595 Å as against the X-ray-determined value of 2.587 Å. The average Eu/Am−Npy distance was found to be 2.695/2.705 Å, whereas the Eu/Am−Namine distance was found to be 2.695 Å. The average Eu/Am−ONO3 bond length was found to be 2.580/ 2.612 Å. Recently,29 the average Eu/Am−Npy and Eu/Am− Namine distance for the charged [M(TPEN)(NO3)2]+ species in the gas phase at the M06-2X/6-31g(d)/RECP level of theory was reported to be 2.698/2.675 and 2.788/2.788 Å, respectively. The present results cannot be directly compared with the reported data29 because of the differences in the method of calculation as well as charge. It is noted that the various M−L bond distances of the Am ion were little longer than that of the Eu ion, indicating its more covalent nature.39 It is worth mentioning that the M−N bond distances using COMS-RS are found to be shorter than that using COSMO, whereas the M−O bond distances using COSMO-RS are found to be longer than that using COSMO (see Figure 2a). 2.1.4. Equilibrium Structure of Complexes of Metal Ions with PPDEN in Water. The optimized minimum-energy structures of the complexes of Eu3+ ion with PPDEN using both COSMO and COSMO-RS are displayed in Figure 3 whereas those of the complexes of Am3+ are shown in Figure S6. The calculated bond length parameters are listed in Table S3. The central metal ion was seen to be coordinated to eight N donor atoms of PPDEN and two donor atoms of the NO3− anion in a bidentate mode, leading to a deca-coordinated

Figure 3. Optimized structures of the complexes of Eu3+ with PPDEN ligand in the presence of nitrate ions using the COSMO and COSMO-RS approaches. 13106

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

Table 1. Calculated Thermodynamical Functions (kcal/mol) for the Complexation Reaction of Hydrated Clusters of La3+, Sm3+, Eu3+, and Am3+ Ions with TPEN Using COSMO and COSMO-RS in Aqueous Solution (ε = 80) 3+

complexation reaction

ΔU

ΔH

ΔSa

ΔGcomplex



−43.8/−38.0 −46.2/−37.2 −50.7/−38.9 −52.3/−41.7

−40.8/−35.0 −43.3/−34.2 −47.8/−35.9 −49.4/−38.8

0.141/0.14 0.147/0.149 0.147/0.15 0.141/0.147

−83.1/−77.1 −87.1/−78.7 −91.7/−80.8 −91.4/−82.6

[La−(H2O)9] + TPEN + 3NO3 = La−TPEN−(NO3)3 + 9H2O [Sm−(H2O)9]3+ + TPEN + 3NO3− = Sm−TPEN−(NO3)3 + 9H2O [Eu−(H2O)9]3+ + TPEN + 3NO3− = Eu−TPEN−(NO3)3 + 9H2O [Am−(H2O)9]3+ + TPEN + 3NO3− = Am−TPEN−(NO3)3 + 9H2O a

cal/(mol·K).

Table 2. Calculated Thermodynamical Functions (kcal/mol) for the Complexation Reaction of Hydrated Clusters of Eu3+ and Am3+ Ions with PPDEN Using COSMO and COSMO-RS in Aqueous Solution(ε = 80) 3+

complexation reaction

ΔU

ΔH

ΔSa

ΔGcomplex

ΔΔGAm−Eu



−34.9/−23.7 −36.0/−22.0

−31.9/−20.7 −33.1/−19.0

0.149/0.183 0.140/0.173

−76.6/−66.6 −74.8/−61.8

1.8/4.7

[Eu−(H2O)9] + PPDEN + 3NO3 = Eu−PPDEN−(NO3)3 + 9H2O [Am−(H2O)9]3+ + PPDEN + 3NO3− = Am−PPDEN−(NO3)3 + 9H2O a

cal/(mol·K).

CCSD,43 MP2,32,44 BP86,32 and earlier-reported B3LYP.32 For the Eu3+ ion, the value ΔGhyd by COSMO-RS is better than that reported from MP2,32,44 BP86,32 and B3LYP.34 The calculated ΔGhyd of Am3+ with COSMO and COSMO-RS was found to be −779.39 and −793.57 kcal/mol, respectively. Unfortunately, there are no experimental results available for ΔGhyd of the Am3+ ion. Therefore, the value was compared with the reported theoretical results. Here, the value calculated using COSMO-RS is in good agreement with the value reported from CCSD45 and quite better than the value reported using BP86,46 B3LYP,47 and semiempirically.47 Finally, the results shown in Table S5 indicate that our computed (B3LYP) values of ΔΔGhyd(ΔGhyd,Eu − ΔGhyd,Am) (−21.6 kcal/mol) using COSMO-RS is in excellent agreement with the recently reported results (−21.13 kcal/mol).20 From the overall analysis, it is seen that COSMO-RS in combination with B3LYP performs better than COSMO and the other reported theoretical results. This further assures the acceptance of the adopted computational methodology here and is thus used in the calculation of the complexation free energy, which is used for the analysis of the stability constant. 2.2.2. Enthalpy and Free Energy of Metal−Ligand Complexation in Water. After obtaining the reasonable structures, we deal with the thermodynamic functions to evaluate the stability of the metal ion−ligand complexes in aqueous solution. The relative stability of the complexes will help in understanding the ability of the extractants to be used in the extraction experiments. Earlier, the stability of Ln/An with TPEN has been reported experimentally by measuring the enthalpy and free energy in aqueous solution. Therefore, the thermodynamical enthalpy and free energy of the complexes in aqueous solution have been determined using both COSMO and COSMO-RS. 2.2.3. M3+−TPEN System. The enthalpies of the complexation reactions for La3+, Sm3+, Eu3+, and Am3+ with TPEN in aqueous solution using both the COSMO and COSMO-RS approaches are presented in Table 1. The complexation enthalpy was found to be negative, indicating the exothermic nature of the reaction as observed in experiments.48 The point to be noted is that the complexation enthalpy calculated using the COSMO approach follows the experimental trend: Am3+ > Sm3+ > La3+, whereas the COSMO-RS approach predicts the trend as Am3+ > La3+ > Sm3+, which does not match the experimental selectivity.

The free energy of complexation for La3+, Sm3+, Eu3+, and Am3+ with TPEN in aqueous solution using both the COSMO and COSMO-RS approaches is presented in Table 1. In the case of both the COSMO and COSMO-RS approaches, the calculated value of the free energy of complexation follows the trend: Am3+ > Sm3+ > La3+, as predicted in the experiment. Overall, the combination of B3LYP and COSMO-RS is thus confirmed as an appropriate computational protocol for determining the complexation stability of Ln/An with TPEN by calculating the free energy of complexation. 2.2.4. M−(NO3)3−PPDEN System. After the above successful demonstration for establishing the experimental complexation stability, our next aim is to test the complexation stability for the higher derivative of TPEN, that is, PPDEN having a higher denticity of 8 “N” donor atoms. The complexation enthalpy of the Eu3+ and Am3+ ions with PPDEN using both COSMO and COSMO-RS is displayed in Table 2. From the table, it is seen that the calculated value of complexation enthalpy of the Am3+ ion with PPDEN using the COSMO solvation model is slightly higher than that of the Eu3+ ion. After the inclusion of positive entropy contribution, the free energy of complexation becomes more exergonic, but the value is seen to be higher for the Eu3+ ion over the Am3+ ion. A similar trend was also noticed with TPEN using COSMO. Further, the complexation enthalpy was evaluated using the COSMO-RS solvation model (Table 2). The complexation enthalpy was exergonic, which after the inclusion of positive entropy contribution becomes more exergonic. Here also, the complexation free energy for the Eu3+ ion was higher than that of the Am3+ ion, which is reversed compared to that observed toward TPEN using COSMO-RS. Furthermore, the calculated free energy with PPDEN is reduced compared to that with TPEN, though PPDEN has a greater denticity (8) than TPEN (6). Therefore, it can be stated that the increased denticity does not always lead to a higher interaction/free energy. A similar reduction in interaction energy with increasing denticity for rare earths with diglycolamide was earlier reported as a manifestation of ligand strain energy.49 In view of this, ligand strain energies (ΔEstrain) were calculated for both TPEN and PPDEN and are presented in Figure 4. The ΔEstrain value was calculated by taking the difference of the total energy of the free ligand and that of the same configuration as in the complex, but after removing the metal ion and anion(s), followed by a single point energy calculation using the COSMO-RS solvation model. The calculated value of ΔEstrain 13107

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

The complexation and the nature of the M3+−ligand bonding can be concluded from the partial atomic charge on the metal ions and from the electron population in the s, d, and f orbitals of the metal ions in the complexes. The electron populations in the atomic orbitals of the TPEN and PPDEN ligands and their complexes with the Am3+ and Eu3+ cations in water are presented in Table S6. The covalent nature of bonding is characterized by the amplification of the electron population in the s, d, and f orbitals of the metal ions subsequent to complexation. There is hardly any difference in the electron population of the metal ion complexes obtained from the COSMO and COSMO-RS models. The LUMO−HOMO energy gaps (LHGs) of TPEN and hydrated metal ions were determined to get a better disclosure of the molecular-level interactions, and the computed values are presented in Table 4. Figure 5 depicts the distribution of the highest occupied molecular orbitals (HOMO) and the lowest occupied molecular orbitals (LUMO) of TPEN in water and NB. The identical nature of HOMO and LUMO of TPEN in water and NB leads to close values of LHGs, indicating weak solvent dependency. The high values of ELUMO−HOMO, χ, and η indicate the hard nature of metal ions. Though the hydrated Eu3+ ion was found to be harder than that of the hydrated Am3+ ion, the interaction energy of the Am3+ ion with TPEN was found to be higher over the Eu3+ ion. Further, the distribution of HOMO and LUMO of the complexes of Eu and Am with TPEN in water and NB is depicted in Figure 6 to realize a better understanding about the type of bonding. From the figure, it is seen that though the distribution of HOMO and LUMO of the complexes of Eu in water and NB are found to be similar, their LHGs are different. The LHG in water is found to be quite higher than that in NB, whereas in case of Am, though the distribution of LUMO is identical in water and NB, the distribution of HOMO is quite dissimilar. The HOMO in water is widely distributed over the ions and ligands, whereas in NB, it is over the nitrate ion only, as observed for Eu. The point to be noted is that for all the cases, the same isosurface value of 0.011 (e·bohr−3) has been used. It is worth mentioning that the distribution of HOMO and LUMO of free TPEN in water and NB is very similar, which leads to a very close value of LHG, whereas in the complexes, the LHG is substantial, indicating the role of the metal orbital making the difference. Further, the distribution of LUMO and HOMO over the metal ion confirms the participation of the f orbital of the metal ions. The extent of interaction of Eu and Am with TPEN is also reflected from the different values of LHG in water and NB. The LHG for the hydrated Am ion is smaller than that for the hydrated Eu ion, indicating that the Am ion is softer than the Eu ion, which perhaps resulted in a higher interaction energy for Am over the Eu ion. Furthermore, the amount of charge transfer, ΔN, was also determined for the metal ion−ligand acceptor donor interaction, and the calculated values are given in Table 4. It

Figure 4. Structures of free TPEN and PPDEN (left) and those after the removal of the metal ion and anion (right) from the respective complexes. The internitrogen distances (in Å) are shown in red dotted color.

for TPEN (25.88 kcal/mol) was found to be much smaller than that of PPDEN (61.23 kcal/mol) and thus suggesting reduced interaction energy compared to TPEN. The higher ΔEstrain for PPDEN compared to TPEN is understandable from the internitrogen distance in the free ligand compared to the organized ligand in the metal ion complex (Figure 4). The internitrogen distance in free PPDEN is much higher than that in TPEN (see Figure 4), and hence ΔEstrain is also much higher than that for TPEN. Also, the cavity of PPDEN was seen to be larger than that of TPEN, as revealed from Figure S9, which resulted in a longer metal-ion−ligand bond distance for PPDEN than TPEN, and hence a smaller interaction energy was observed for PPDEN. It is worth noting that not only the free energy is reduced, but selectivity is also reversed. The difference in strain energies, ΔΔEstrain (=ΔEstrain,Eu − ΔEstrain,Am), for TPEN (2.29 kcal/mol) was found to be reduced to 1.73 kcal/mol for PPDEN, indicating the possible role of strain energy in reversing the TPEN/PPDEN selectivity. In the next section, the bonding analysis has been performed to find out the other probable reasons for the reduction in the free energy of complexation from TPEN to PPDEN. 2.3. Bonding Analysis. To obtain better molecular insights in the nature of bonding of Eu/Am with TPEN and PPDEN, natural population analysis (NPA)50,51 was carried out on various chemical species using both COSMO and COSMO-RS. 2.3.1. Metal Ion−TPEN Complexes. The residual charge calculated by performing NPA on the TPEN complex of Eu ion was found to be the lowest, whereas it was the highest with the Sm ion, though the difference is very minute (Table 3).

Table 3. Calculated Charge and Orbital Population of the Complexes of the Sm3+, Eu3+, and Am3+ Ions with TPEN Using NBO Analysis Using COSMO and COSMO-RS in Aqueous Solution (ε = 80) complex species

charge on metal ion (esu)

n(s)

n(p)

n(d)

n(f)

n(g)

TPEN−Sm(NO3)3 TPEN−Eu(NO3)3 TPEN−Am(NO3)3

1.723/1.732 1.699/1.705 1.725/1.730

4.19/4.19 4.19/4.19 4.20/4.10

11.99/11.99 11.99/11.99 11.99/11.99

10.98/10.97 10.98/10.96 10.92/10.91

5.10/5.10 6.12/6.13 6.14/6.14

0.0007/0.0006 0.0006/0.006 0.0008/0.0007

13108

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

Table 4. Calculated Quantum Chemical Descriptors Using COSMO and COSMO-RS of Hydrated La3+, Sm3+, Eu3+, and Am3+ Ions and TPEN in Aqueous Solution (ε = 80) [La−(H2O)9]3+ [Sm−(H2O)9]3+ [Eu−(H2O)9]3+ [Am−(H2O)9]3+ TPEN

EHUMO (eV)

ELUMO (eV)

ELUMO−HOMO (eV)

χ (eV)

η (eV)

ΔN

−9.89/−9.22 −9.86/−9.18 −9.85/−9.16 −7.86/−7.18 −5.66/−5.93

−1.11/−0.15 −4.98/−4.28 −5.96/−5.26 −4.07/−3.42 −0.94/−1.17

8.77/9.07 4.88/4.89 3.88/3.92 3.79/3.75 4.71/4.75

5.50/4.69 7.42/6.73 7.91/7.22 5.97/5.30 3.30/3.55

4.38/4.53 2.44/2.44 1.94/1.96 1.89/1.87 2.35/2.37

0.16/0.08 0.42/0.32 0.53/0.42 0.31/0.20

Figure 5. Distribution of HOMO and LUMO of TPEN and PPDEN in water and NB.

Figure 6. Distribution of the HOMO and LUMO of the complexes of Eu and Am with TPEN in water and NB.

Table 5. Calculated Charge on N Donor in Free TPEN and Complexes Using COSMO and COSMO-RS Phases in Aqueous Solution charge system TPEN TPEN−Sm(NO3)3 TPEN−Eu(NO3)3 TPEN−Am(NO3)3

Npy −0.443/−0.468 −0.480/−0.484 −0.493/−0.496 −0.500/−0.503

−0.441/−0.467 −0.484/−0.491 −0.482/−0.488 −0.495/−0.501

Neth‑amine −0.472/−0.509 −0.495/−0.497 −0.476/−0.485 −0.487/−0.493

−0.448/−0.463 −0.480/−0.487 −0.479/−0.485 −0.487/−0.490

−0.391/−0.394 −0.432/−0.435 −0.424/−0.423 −0.437/−0.439

−0.399/−0.400 −0.429/−0.431 −0.428/−0.430 −0.440/−0.442

obtained using COSMO. In the case of COSMO-RS, though the charge transfer, ΔN, was estimated to be higher with Eu3+ ion than that with Am3+ ion, the value of (ΔU) was found to be higher with Am3+ ion over that with Eu3+ ion and hence cannot be correlated. Further, the charge on the donor “N” atom of pyridine and ethylene amine is calculated employing

is worth noting that the calculated large ion−ligand interaction energy (ΔU) as well as the higher interaction of the Eu3+ ion over the Am3+ ion cannot be correlated with the higher amount of charge transfer, ΔN, with the Eu3+ ion than with the Am3+ ion. The point to be noted is that the value of ΔN predicted using COSMO-RS is slightly smaller than that 13109

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

Table 6. Calculated Charge and Orbital Population of Complexes of Eu3+ and Am3+ Ions with PPDEN Using NBO Analysis Using COSMO and COSMO-RS Phases in Aqueous Solution complex species

charge on metal ion (esu)

n(s)

n(p)

n(d)

n(f)

n(g)

PPDEN−Eu(NO3)3 PPDEN−Am(NO3)3

1.779/1.776 1.803/1.819

4.17/4.19 4.19/4.19

11.99/11.99 11.99/11.99

10.87/10.86 10.87/10.85

6.16/6.17 6.13/6.13

0.0004/0.004 0.0005/0.0005

Table 7. Calculated Quantum Chemical Descriptors Using COSMO and COSMO-RS of the Complexes of Eu3+ and Am3+ Ions with PPDEN in Aqueous Solution [Eu−(H2O)9]3+ [Am−(H2O)9]3+ PPDEN

EHUMO (eV)

ELUMO (eV)

ELUMO−HOMO (eV)

χ (eV)

η (eV)

ΔN

−9.85/−9.16 −7.86/−7.18 −5.39/−5.76

−5.96/−5.26 −4.07/−3.42 −0.85/−1.22

3.88/3.92 3.79/3.75 4.53/4.53

7.91/7.22 5.97/5.30 3.12/3.49

1.94/1.96 1.89/1.87 2.26/2.26

0.56/0.44 0.34/0.21

Figure 7. Distribution of HOMO and LUMO of the complexes of Eu and Am with PPDEN in water and NB.

The LHG of PPDEN was determined to obtain a better understanding about the molecular-level interaction, and the calculated values are given in Table 7. Figure 7 depicts the dispersal of HOMO and LUMO of PPDEN in water and NB. The values of LHG of PPDEN are similar in water and NB because of the identical nature of HOMO and LUMO. From the figure, it is seen that though the distribution of HOMO and LUMO of PPDEN in water and NB is quite identical, LHG has been reduced compared to that of TPEN. Further, the distribution of HOMO and LUMO of the complexes of Eu and Am with PPDEN in water and NB is depicted in Figure 7 to have a better understanding about the type of bonding. From the figure, it is noticed that the nature of the HOMO and LUMO of the complexes of Eu in water and NB is not only found to be comparable, but their LHGs are almost matching as noticed in the case of free TPEN, whereas in the case of Am, though the distribution of LUMO and HOMO is identical in water and NB, their LHGs are dissimilar. The LHG in water is found to be quite higher than that in NB. It is worth noting that the distribution of HOMO and LUMO of free PPDEN in water and NB is very similar, leading to equal LHGs, whereas in the complexes, the LHG has been substantially reduced indicating the role of metal orbital which makes the difference. Further, the distribution of LUMO and HOMO over the metal ion confirms the participation of the f orbital of metal ions. The extent of interaction of Eu and Am with PPDEN is also reflected from the different values of LHG in water and NB. The high values of ELUMO−HOMO, χ, and η indicate the hard nature of all the metal ions studied here. Though the hydrated Eu3+ ion was found to be harder than the hydrated Am3+ ion, the interaction

both COSMO and COSMO-RS, and the calculated results are presented in Table 5. From the table, it is noticed that the partial charge on the N atoms of the pyridine moieties, higher than that on the N atoms of the ethylene amine chain, indicates the stronger complexation of the metal cations by the Npy than that by the Neth‑amine atoms. The partial charge on the “N” donor in free TPEN as well as in the metal ion complexes using COSMO-RS was found to be slightly higher than that using COSMO. The augmentation of charge on the donor “N” atoms after complexation indicates a strong ion−dipole interaction. Npy and Neth‑amine atoms in the coordinated TPEN (Table 5) molecules are more negative than those in the respective free molecules. This is not unusual and has earlier been reported for the terdendate nitrogen planar ligands52 and may be attributed to the withdrawal of electron density from the less negative carbon atoms of the ligands to the more negative N atoms (see Table S6). Note that the sum of the partial charges on all the atoms is zero. 2.3.2. Metal Ion−PPDEN Complexes. The residual charge on the PPDEN complex of Eu ion was found to be smaller than that of the Am ion, though the difference is very small (Table 6). The residual charge on the metal ion in the COSMO approach is very close to that in the COSMO-RS approach. The residual charge on both the Eu and Am ions in the complexes of octadentate PPDEN was found to be increased compared to that of the hexa-dentate TPEN, indicating a reduced interaction with PPDEN over TPEN. There is hardly any difference in the electron population of the metal ion complexes obtained from COSMO and COSMORS. 13110

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

Table 8. Calculated Charge on N Donor in Free PPDEN and Complexes Using COSMO and COSMO-RS in Aqueous Solution charge system

Npy

Neth‑amine

PPDEN PPDEN−Eu(NO3)3 PPDEN−Am(NO3)3

−0.469, −0.467, −0.468, −0.429, −0.450/−0.508, −0.504, −0.502, −0.467, −0.466 −0.477, −0.470, −0.499, −0.494, −0.464/−0.482, −0.469, −0.495, −0.465, −0.503 −0.486, −0.472, −0.508, −0.465, −0.507/−0.475, −0.481, −0.513, −0.462, −0.520

−0.415, −0.395, −0.442/−0.410, −0.392, −0.439 −0.442, −0.431, −0.418/−0.441, −0.430, −0.415 −0.454, −0.441, −0.436/−0.454, −0.451, −0.438

energy of the Eu3+ ion with PPDEN was found to be higher than that of the Am3+ ion when the COSMO solvation model is employed. Furthermore, the amount of charge transfer, ΔN, was also determined for the metal ion−ligand acceptor donor interaction, and the calculated values are given in Table 7. It is worth noting that the calculated large ion−ligand interaction energy can be well-correlated with the higher amount of charge transfer, ΔN, but could not be correlated with the higher interaction of the Am3+ ion over that of the Eu3+ ion. It is noted that the value of ΔN predicted using COSMO-RS is slightly smaller than that obtained using COSMO solvation. However, in case of COSMO-RS, the estimated charge transfer, ΔN, as well as the interaction energy, ΔU, was shown to be higher with the Eu3+ ion than that with the Am3+ ion and is well-correlated. Further, the charge on the donor “N” atom of pyridine and ethylene amine in PPDEN and the metal ion complex is calculated employing both COSMO and COSMO-RS, and the calculated results are presented in Table 8. From the table, it is noticed that the partial charge on the donor “N” atom within the pyridine was found to be higher than that with the ethylene “N” donor, indicating a strong interaction with the pyridine “N” over the ethylene “N” as observed with TPEN. The partial charge on the “N” donor in free PPDEN as well as in the metal ion complexes using COSMO-RS was found to be close to that using COSMO solvation. As in TPEN, the Npy and Neth‑amine atoms in the coordinated PPDEN (Table 8) molecules are more negative than the respective free ligand, which might be due to the withdrawal of electron density from the less negative carbon atoms of the ligands to the more negative N atoms (see Table S6). Overall, from the population and LUMO−HOMO analyses, the interaction of PPDEN is found to be smaller than that of TPEN, and also different ionic, covalent, and ion−dipole type bondings were revealed. From the analysis, TPEN was shown to be harder than PPDEN. The reactivity of the TPEN and PPDEN ligands because of their different denticities can also be well-depicted by the molecular electrostatic potential (MEP). The MEP is very useful as an indicator of the sites of a ligand molecule to which an approaching ionic species is complexed. In view of this, we have calculated the MEPs of TPEN and PPDEN at the B3LYP/TZP level of theory using the ADF suite of package53,54 and depicted in Figure 8. The MEP clearly reveals the different charge distributions of TPEN and PPDEN because of their different denticities. Moreover, the bonding character of the complexes of Am3+ and Eu3+ with TPEN was studied by evaluating the energy levels of the MOs at the B3LYP/TZ2P/ZORA level of theory. The MO diagrams of LUMO, LUMO + 1, HOMO, and HOMO − 1 for the α-spin valence of the metal−ligand bonding for the [TPEN−M]3+ complexes (an isosurface value of 0.017 a.u. was used) are shown in Figure 9. From the analysis of MOs, it is seen that there is no participation of the f

Figure 8. MEPs of TPEN and PPDEN at the B3LYP/TZ2P level of theory.

orbital of Eu in HOMO, whereas the participation of the f orbital of Am in HOMO was observed.

Figure 9. Calculated energy levels (eV) of the α-spin valence MOs of the [TPEN−M]3+ complexes using the B3LYP/TZ2P/ZORA level of theory.

2.4. Selectivity of Eu3+ and Am3+ Toward TPEN/ PPDEN in Biphasic (Water−NB) Extraction. After establishing the complexation stability and bonding mechanism, our next attempt is to exploit this behavior for preferential partitioning of Eu and Am with TPEN/PPDEN using liquid−liquid extraction. In view of the higher complexation free energy of Am3+ with TPEN over Ln in aqueous solution, it will be worthwhile to investigate whether TPEN can be used for the partitioning of actinides (Am) from Ln (Eu) in biphasic solvent extraction experiments. For this, the biphasic extraction reaction can be modeled as [M − (H 2O)9 ]3 + (aq) + L(nitrobenzene) + 3NO3−(aq) = [M − L − (NO3)2 ]NO3(nitrobenzene) + 9H 2O(aq) (L = TPEN/PPDEN)

(1)

The free energy of extraction, ΔGext, for the above complexation reaction has been evaluated using the earlier reported method,55−60 and the values are presented in Table 9. In Table 9, the value of ΔG was determined using the 13111

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

Table 9. Calculated Thermodynamical Parameters (kcal/mol) for the Extraction Reaction (W/NB) of Hydrated Clusters of Eu3+ and Am3+ Ions with TPEN and PPDEN Using COSMO Solvation Model ΔU

ΔH

ΔSa

ΔGext

ΔΔGAmEu

−49.0/−25.5 −50.6/−30.2 −32.3/−1.8 −33.1/−0.88

−46.0/−22.5 −47.6/−27.2 −29.3/1.1 −30.2/2.0

0.146/0.148 0.139/0.148 0.156/0.153 0.146/0.138

−89.6/−66.9 −89.3/−71.4 −75.9/−44.5 −73.8/−39.1

0.4/−4.5

complexation reaction 3+



[Eu−(H2O)9] + TPEN + 3NO3 = Eu−TPEN−(NO3)3 + 9H2O [Am−(H2O)9]3+ + TPEN + 3NO3− = Am−TPEN−(NO3)3 + 9H2O [Eu−(H2O)9]3+ + PPDEN + 3NO3− = Eu−PPDEN−(NO3)3 + 9H2O [Am−(H2O)9]3+ + PPDEN + 3NO3− = Am−PPDEN−(NO3)3 + 9H2O

2.0/5.42

a

cal/(mol·K).

calculation at a higher level of theory or by synthesizing a new PPDEN ligand followed by experimental evaluation. It is noted that a similar computational study for the complexation of Eu and Am with TPEN using DFT/COSMO was recently reported by Huang et al.29 with the theoretical selectivity of Am over Eu as observed in the present study at the B3LYP/ TZVP level of theory in conjunction with COSMO-RS.

COSMO solvation model. Further, the selectivity between the Am3+ ion and the Eu3+ ion is determined using the difference in the free energy of extraction, ΔΔG = ΔGext,Am − ΔGext,Eu. The negative ΔG value indicates that the ligand is Amselective and positive ΔG value means Eu-selective. In case of TPEN, the value of ΔGext for Eu3+ ion was found to be slightly higher than that of Am3+ ion by 0.38 kcal/mol, and with PPDEN, the value of ΔGext was found to be favorable by 2.05 kcal/mol for Eu3+ ion over Am3+ ion. From the calculated values of ΔGext, both TPEN and PPDEN were shown to be Eu3+ ion-selective over Am3+ ion when the COSMO solvation model was used, which contradicts the preferential selectivity of the Eu3+ ion over the Am3+ ion, as observed in the experiments (DAm = 2.0 and DEu = 0.0099).7,8 The calculated value of ΔGext using COSMO-RS is also presented in Table 9. From the calculated values of ΔGext, it is seen that with TPEN, the value of ΔGext for the Am3+ ion was found to be higher than that of the Eu3+ ion by 4.5 kcal/mol, indicating that TPEN is selective for Am3+ ion over Eu3+ ion, as observed in the experiments (DAm = 2.0 and DEu = 0.0099),7,8 whereas with PPDEN, the value of ΔGext was found to be favorable by 5.42 kcal/mol for Eu3+ ion over Am3+ ion, predicting PPDEN is Eu3+ ion-selective. From the value of ΔGext, TPEN was found to be Am3+ ion-selective over Eu3+ ion, and PPDEN was shown to be Eu3+ ion-selective when the COSMO-RS solvation model was employed. Note that the calculated value of ΔΔGAm−Eu (−4.5 kcal/mol) for TPEN in NB using COSMO-RS was found to be very close to the earlier reported value of −5.1 kcal/mol29 at the M06-2X/6-31g(d)/RECP level of theory, although the type of complex was [M−TPEN−(NO3)2]. The point to be noted is that the binding energy, ΔU, for both the metal ions with both TPEN and PPDEN was found to be reduced for COSMO-RS over the COSMO model substantially. This might be due to the calculated M−O bond length using COSMO-RS, which is seen to be slightly longer than that using COSMO, as longer the bond length, smaller is the interaction. From the value of ΔU, both TPEN and PPDEN were found to be Am3+-selective using the COSMO model, but Am3+-selective for TPEN and Eu3+-selective for PPDEN using the COSMO-RS model. After the inclusion of entropy, both TPEN and PPDEN were found to be Eu3+selective using the COSMO model, whereas TPEN remains Am3+-selective and PPDEN remains Eu3+-selective using the COSMO-RS model. Anticipating that the COSMO-RS model is more accurate over the COSMO model because of its superior functionality by taking care of the structure of the solvents, one might conclude that TPEN is Am3+-selective, whereas PPDEN is Eu3+-selective, which might be due to the steric denticity of PPDEN than that of TPEN. This contrasting selectivity of TPEN and PPDEN can be exploited for the partitioning of Eu 3+ and Am3+ in the nuclear waste reprocessing. Finally, the predictability of the COSMO-RSenabled DFT calculation can be confirmed by performing the

3. CONCLUSIONS The enthalpy of the complexation reaction for metal ions with TPEN and PPDEN in aqueous solution was found to be negative, indicating the exothermic nature of the reaction as observed in the experiments. The calculated value of the complexation enthalpy of the Eu3+ ion with PPDEN is slightly higher than that of the Am3+ ion. The point to be noted is that the complexation free energy of the Eu3+ ion was higher than that of the Am3+ ion with PPDEN, which is opposite to that with TPEN using COSMO-RS. Furthermore, the calculated free energy with PPDEN is reduced compared to that with TPEN. This may be due to the higher ligand straining for PPDEN than for TPEN. Therefore, it can be concluded that the increased denticity always does not lead to more negative free energy. It is worth mentioning that the free energy with PPDEN is reduced and selectivity is reversed. The partial charge on the N atoms of the pyridine moieties, higher than that on the N atoms of the ethylene amine chain, indicates a stronger complexation of the metal cations by Npy than by the Neth‑amine atoms. The increased partial charge on the donor “N” atoms reveals the strong ion−dipole interaction. The residual charge on the Eu/Am ions in the complexes of PPDEN was higher compared to that of TPEN, indicating a reduced interaction with PPDEN than that with TPEN. Further, the covalent nature of bonding is manifested by the greater electron population in the s, d, and f orbitals of metal ions after complexation. Anticipating that the COSMO-RS model is more accurate over the COSMO model (demonstrated for metal ion hydration free energy with COSMO-RS) because of its superior functionality by taking care of the structure of the solvents, one might conclude that TPEN is Am3+-selective, whereas PPDEN is Eu3+-selective, which might be due to the higher ligand straining with PPDEN than that with TPEN. Thus, the contrasting selectivity of TPEN and PPDEN might be exploited for the effective partitioning of the Eu3+ and Am3+ ions from the high-level nuclear waste. Finally, the predictability of the COSMO-RS-enabled DFT calculations needs further investigation, either by performing the calculation at a higher level of theory or by experiments using PPDEN. Work in this direction is in progress in our computational and experimental molecular engineering laboratory. 13112

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

Figure 10. Sigma potential of H2O and NB used for the optimization of structures, and the calculation of energy and frequency using COSMO-RS.

elements except La3+, Sm3+, Eu3+, and Am3+, for which the values of COSMO radii used were 2.03, 1.93, 1.90, and 1.99 Å, respectively.31,45 The dielectric constants, ε, of 80 and 34.81 were used to represent the water and NB environment, respectively. The computation of the solvation free energy for the metal ions in water is performed using the cluster water explicit solvation model.31,45 The complexation reaction in the aqueous phase was modeled using 1:1 metal/ligand stoichiometry, as observed in the experiments, which is given below32

4. COMPUTATIONAL METHODOLOGY The hydrated clusters of trivalent Ln (La3+, Sm3+, and Eu3+) and actinides (Am3+), TPEN and PPDEN ligands, and their metal nitrate complexes have been optimized in the aqueous and NB phases at the B3LYP level of theory61,62 using the TZVP basis set63,64 as available in the electronic structure calculation program of Turbomole,65−67 that is, C (11s6p1d)/ [5s3p1d], N (11s6p1d)/[5s3p1d], O (11s6p1d)/[5s3p1d], and H (5s1p)/[3s1p] and the SVP basis set, that is, La (7s6p5d)/[6s3p2d], Sm (14s13p10d8f1g)/[10s8p5d4f1g], Eu (14s13p10d8f1g)/[10s8p5d4f1g], and Am (14s13p10d8f1g)/ [10s9p5d4f1g]. In the case of La3+, Sm3+, Eu3+, and Am3+, an ECP core potential was used, where 46 electrons are kept in the core of La68−70 and 28 electrons are kept in the core of Sm70 and Eu,71−73 respectively, whereas the number of core electrons was 60 for the Am71−73 element. For other elements, all-electron basis set was used. The aqueous and NB phase effects were taken care by invoking both the COSMO74 and COSMO-RS75,76 solvation approaches. The COSMO solvation model, which is a simply dielectric continuum model, does not take into account the microscopic configuration of the solvents. To account the microscopic structure of the solvents, a self-consistent module of COSMORS75,76 has been adopted, where the sigma potential of individual solvents can be used. The sigma potential captures the microscopic details of the solvents through the surface charge density and has been computed here by performing the D-COSMO-RS77,78 module as implemented in the Turbomole program. The sigma potentials of water and NB, which are needed for the COSMO-RS calculation, are depicted in Figure 10. The optimization of the complexes with Sm3+, Eu3+, and Am3+ was performed using the doublet and septet spin-state, respectively. There are numerous studies in the literature, where the B3LYP functional has been used successfully for the calculations of the molecular properties of high-spin Eu and Am systems.33,79−83 Hence, B3LYP was selected for the calculations of the present Eu and Am systems. The issue of degeneracy because of the spin is generally checked by comparing the value of ⟨S2⟩ and S(S + 1). In the present chemical system, the value of ⟨S2⟩ is observed to be very close to the S(S + 1) ideal values, which highly indicates that all the chemical structures display negligible spin contamination. The vibrational frequency was calculated using the numerical NumForce module of the Turbomole. The default COSMO radii were used for all the

[M − (H 2O)9 ]3 + (aq) + 3NO3−(aq) + TPEN(aq) = [M − TPEN − (NO3)2 ]NO3(aq) (M = La, Sm, Eu, and Am)

(2)

The zero-point energy corrected complexation energy for the above reaction can be written as ΔEcomplx = E[M − TPEN − (NO3)2 ]NO3(aq) − (E M3+− (H2O)9 (aq) + 3E NO3−(aq) + E TPEN(aq)) (3)

The enthalpy of complexation, ΔHcomplx, is obtained by the following thermal corrections ΔHcomplx = ΔEcomplx + ΔnRT

(4)

Here, Δn is the change in the number of moles between the products and reactants. The free energy for the complexation reaction is determined as ΔGcomplx = ΔHcomplx − T ΔScomplx

(5)

where ΔS is the entropy of complexation and T is the temperature. NPA50,51 was performed to understand the nature of bonding in the metal−ligand complexes. Further, from the LUMO−HOMO analysis, the strength of the M−L interaction can also be evaluated by the amount of change transfer (ΔN),84 which can be calculated by eq 6 as ΔN =

(χM − χL ) {2(ηM + ηL)}

(6)

where χ and η represent the absolute electronegativity and absolute hardness of M or L, respectively, and estimated as χ= 13113

I+A I−A and η = 2 2

(7) DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

partitioning and transmutation scenarios on the radiotoxicity of actinides in radioactive waste. Nucl. Energy 2003, 42, 263−277. (4) Ekberg, C.; Fermvik, A.; Retegan, T.; Skarnemark, G.; Foreman, M. R. S.; Hudson, M. J.; Englund, S.; Nilsson, M. An overview and historical look back at the solvent extraction using nitrogen donor ligands to extract and separate An(III) from Ln(III). Radiochim. Acta 2008, 96, 225−233. (5) Panak, P. J.; Geist, A. Complexation and Extraction of Trivalent Actinides and Lanthanides by Triazinylpyridine N-Donor Ligands. Chem. Rev. 2013, 113, 1199−1236. (6) Whittaker, D. M.; Griffiths, T. L.; Helliwell, M.; Swinburne, A. N.; Natrajan, L. S.; Lewis, F. W.; Harwood, L. M.; Parry, S. A.; Sharrad, C. A. Lanthanide Speciation in Potential SANEX and GANEX Actinide/Lanthanide Separations Using Tetra-N-Donor Extractants. Inorg. Chem. 2013, 52, 3429−3444. (7) Matsumura, T.; Takeshita, K. Extraction Behavior of Am(III) from Eu(III) with Hydrophobic Derivatives ofN,N,N’,N’-tetrakis(2methylpyridyl)ethylenediamine (TPEN). J. Nucl. Sci. Technol. 2006, 43, 824−827. (8) Fugate, G. A.; Takeshita, K.; Matsumura, T. Separation of Americium (III) and Lanthanide (III) Ions using TPEN Analogs. Sep. Sci. Technol. 2008, 43, 2619−2629. (9) Mirvaliev, R.; Watanabe, M.; Matsumura, T.; Tachimori, S.; Takeshita, K. Selective Separation of Am(III) from Ln(III) with a Novel Synergistic Extraction System, N ,N, N, N-tetrakis(2methylpyridyl)-ethylenediamine (TPEN) and Carboxylic Acid in 1Octanol. J. Nucl. Sci. Technol. 2004, 41, 1122−1124. (10) Matsumura, T.; Takeshita, K. Extraction separation of trivalent minor actinides from lanthanides with hydrophobic derivatives of TPEN. Prog. Nucl. Energy 2008, 50, 470−475. (11) Watanabe, M.; Mirvaliev, R.; Tachimori, S.; Takeshita, K.; Nakano, Y.; Morikawa, K.; Chikazawa, T.; Mori, R. Selective Extraction of Americium(III) over Macroscopic Concentration of Lanthanides(III) by Synergistic System of TPEN and D2EHPA in 1Octanol. Solvent Extr. Ion Exch. 2004, 22, 377−390. (12) Weaver, B.; Kappelmann, F. A. Preferential extraction of lanthanides over trivalent actinides by monoacidic organophosphates from carboxylic acids and from mixtures of carboxylic and aminopolyacetic acids. J. Inorg. Nucl. Chem. 1968, 30, 263−272. (13) Benay, G.; Schurhammer, R.; Wipff, G. Basicity, complexation ability and interfacial behavior of BTBPs: a simulation study. Phys. Chem. Chem. Phys. 2011, 13, 2922−2934. (14) Singha Deb, A. K.; Ali, S. M.; Shenoy, K. T.; Ghosh, S. K. Adsorption of Eu3+and Am3+ion towards hard donor-based diglycolamic acid-functionalised carbon nanotubes: density functional theory guided experimental verification. Mol. Simul. 2015, 41, 490− 503. (15) Cao, X.; Heidelberg, D.; Ciupka, J.; Dolg, M. First-Principles Study of the Separation of Am(III)/Cm(III) From Eu(III) With Cyanex301. Inorg. Chem 2010, 49, 10307−10315. Roy, L. E.; Bridges, N. J.; Martin, L. R. Theoretical insights into covalency driven f element separations. Dalton Transaction 2013, 42, 2636−2642. (16) Wang, C.-Z.; Lan, J.-H.; Wu, Q.-Y.; Zhao, Y.-L.; Wang, X.-K.; Chai, Z.-F.; Shi, W.-Q. Density functional theory investigations of the trivalent lanthanide and actinide extraction complexes with diglycolamides. Dalton Trans. 2014, 43, 8713−8720. (17) Narbutt, J.; Wodyński, A.; Pecul, M. The selectivity of diglycolamide (TODGA) and bis-triazine-bipyridine (BTBP) ligands in actinide/lanthanide complexation and solvent extraction separationa theoretical approach. Dalton Trans. 2015, 44, 2657−2666. (18) Roy, L. E.; Martin, L. R. Theoretical prediction of coordination environments and stability constants of lanthanum lactate complexes in solution. Dalton Trans. 2016, 45, 15517−15522. (19) Wu, H.; Wu, Q.-Y.; Wang, C.-Z.; Lan, J.-H.; Liu, Z.-R.; Chai, Z.-F.; Shi, W.-Q. Theoretical insights into the separation of Am(III) over Eu(III) with PhenBHPPA. Dalton Trans. 2015, 44, 16737− 16745.

where I and A correspond to the ionization potential and electron affinity, respectively, and can be obtained using Koopmans’ theorem85 as I = − E HOMO A = −E LUMO

(8)

where EHOMO and ELUMO are the energies of the highest occupied and lowest unoccupied molecular orbitals of M or L, respectively, and can be determined from the DFT calculation.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b00933. Calculated structural parameters (Å) of the hydrated clusters of La3+, Sm3+, Eu3+, and Am3+ ions; their complexes with TPEN and PPDEN in water and NB at the B3LYP/TZVP level of theory using COSMO and COSMO-RS; calculated hydration free energy of the hydrated clusters of La3+, Sm3+, Eu3+, and Am3+ ions; optimized hydrated clusters of Eu 3+ (H 2 O) 9 and Am3+(H2O)9 using COSMO and COSMO-RS; optimized TPEN and PPDEN using COSMO; optimized structure of the complex of Sm3+, Eu3+, and Am3+ with TPEN ligand in the presence of nitrate ion using COSMO and COSMO-RS; optimized structure of the complex of Am3+ with PPDEN ligand in the presence of nitrate ion using COSMO and COSMO-RS; optimized structure of the complex of Eu3+ and Am3+ with the TPEN and PPDEN ligands in the presence of nitrate ion using COSMO and COSMO-RS approaches (NB); cavity of TPEN and PPDEN after the removal of metal ions from the complexes; atomic populations from the total density in aqueous phase using COSMO-RS; and Cartesian coordinates of the optimized minimum-energy structures of the hydrated clusters of La3+, Sm3+, Eu3+, and Am3+ ions, their complexes with TPEN and PPDEN in water and NB using COSMO and COSMO-RS(PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Phone: +91-22-25591992. ORCID

Sk. Musharaf Ali: 0000-0003-0457-0580 Notes

The author declares no competing financial interest.



ACKNOWLEDGMENTS Computer Division BARC is acknowledged for providing ANUPAM Supercomputing facility. We sincerely thank Dr. Sadhana Mohan, Associate Director, ChEG, BARC and Shri K.T. Shenoy, Head, Chemical Engineering Division, BARC for their continuous support and encouragement.



REFERENCES

(1) Mathur, J. N.; Murali, M. S.; Nash, K. L. Actinide partitioning-A Review. Solvent Extr. Ion Exch. 2001, 19, 357−390. (2) Nash, K. L.; Madic, C.; Mathur, J. N. The Chemistry of the Actinide and Transactinide Elements; Morss, L. R., Edelstein, N., Fuger, J., Katz, J. J., Eds.; Springer Netherlands, 2006; Vol. 4, pp 2622−2798. (3) Magill, J.; Berthou, V.; Haas, D.; Galy, J.; Schenkel, R.; Wiese, H.-W.; Heusener, G.; Tommasi, J.; Youinou, G. Impact limits of 13114

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

(20) Bryantsev, V. S.; Hay, B. P. Theoretical prediction of Am(III)/ Eu(III) selectivity to aid the design of actinide-lanthanide separation agents. Dalton Trans. 2015, 44, 7935−7942. (21) Fryer-Kanssen, I.; Austin, J.; Kerridge, A. Topological Study of Bonding in Aquo and Bis(triazinyl)pyridine Complexes of Trivalent Lanthanides and Actinides: Does Covalency Imply Stability? Inorg. Chem. 2016, 55, 10034−10042. (22) Lan, J.-H.; Shi, W.-Q.; Yuan, L.-Y.; Zhao, Y.-L.; Li, J.; Chai, Z.F. Trivalent Actinide and Lanthanide Separations by Tetradentate Nitrogen Ligands: A Quantum Chemistry Study. Inorg. Chem. 2011, 50, 9230−9237. (23) Lan, J.-H.; Shi, W.-Q.; Yuan, L.-Y.; Feng, Y.-X.; Zhao, Y.-L.; Chai, Z.-F. Thermodynamic study on the complexation of Am (III) and Eu (III) with tetradentate nitrogen ligands: a probe of complex species and reactions in aqueous solution. J. Phys. Chem. A 2011, 116, 504−511. (24) Lan, J.-H.; Shi, W.-Q.; Yuan, L.-Y.; Li, J.; Zhao, Y.-L.; Chai, Z.F. Recent advances in computational modeling and simulations on the An(III)/Ln(III) separation process. Coord. Chem. Rev. 2012, 256, 1406−1417. (25) Narbutt, J.; Oziminski, W. P. Selectivity of bis-triazinyl bipyridine ligands for americium(III) in Am/Eu separation by solvent extraction. Part 1. Quantum mechanical study on the structures of BTBP complexes and on the energy of the separation. Dalton Trans. 2012, 41, 14416. (26) Xiao, C.-L.; Wang, C.-Z.; Lan, J.-H.; Yuan, L.-Y.; Zhao, Y.-L.; Chai, Z.-F.; Shi, W.-Q. Selective separation of Am(III) from Eu(III) by 2,9-Bis(dialkyl-1,2,4-triazin-3-yl)-1,10-phenanthrolines: a relativistic quantum chemistry study. Radiochim. Acta 2014, 102, 875. (27) Ali, S. M.; Pahan, S.; Bhattacharyya, A.; Mohapatra, P. K. Complexation thermodynamics of diglycolamide with f-elements: solvent extraction and density functional theory analysis. Phys. Chem. Chem. Phys. 2016, 18, 9816−9828. (28) Kaneko, M.; Watanabe, M.; Matsumura, T. The Separation Mechanism of Am(iii) From Eu(iii) by Diglycolamide and Nitrilotriacetamide Extraction Reagents Using DFT Calculations. Dalton Trans. 2016, 45, 17530−17537. (29) Huang, P.-W.; Wang, C.-Z.; Wu, Q.-Y.; Lan, J.-H.; Song, G.; Chai, Z.-F.; Shi, W.-Q. Uncovering the impact of capsule shaped amine-type ligands on Am(III)/Eu(III) separation. Phys. Chem. Chem. Phys. 2018, 20, 1030−1038. (30) Ohta, A.; Kagi, H.; Tsuno, H.; Nomura, M.; Kawabe, I. Influence of multi-electron excitation on EXAFS spectroscopy of trivalent rare-earth ions and elucidation of change in hydration number through the series. Am. Mineral. 2008, 93, 1384−1392. (31) Persson, I.; D’Angelo, P.; De Panfilis, S.; Sandström, M.; Eriksson, L. Hydration of Lanthanoid(III) Ions in Aqueous Solution and Crystalline Hydrates Studied by EXAFS Spectroscopy and Crystallography: The Myth of the “Gadolinium Break”. Chem.Eur. J. 2008, 14, 3056−3066. (32) Ciupka, J.; Cao-Dolg, X.; Wiebke, J.; Dolg, M. Computational study of lanthanide(III) hydration. Phys. Chem. Chem. Phys. 2010, 12, 13215−13223. (33) Morss, L. R.; Rogers, R. D. Syntheses and crystal structures of [M(NO3)2(tpen)][NO3]·3H2O (M=La, Tb), rare earth complexes with strong M-N bonds. Inorg. Chim. Acta 1997, 255, 193−197. (34) Sadhu, B.; Sundararajan, M.; Bandyopadhyay, T. Efficient Separation of Europium Over Americium Using Cucurbit-[5]-uril Supramolecule: A Relativistic DFT Based Investigation. Inorg. Chem. 2016, 55, 598−609. (35) Allen, P. G.; Bucher, J. J.; Shuh, D. K.; Edelstein, N. M.; Craig, I. Coordination chemistry of trivalent lanthanide and actinide ions in dilute and concentrated chloride solutions. Inorg. Chem. 2000, 39, 595−601. (36) Stumpf, T.; Hennig, C.; Bauer, A.; Denecke, M. A.; Fanghänel, T. An EXAFS and TRLFS study of the sorption of trivalent actinides onto smectite and kaolinite. Radiochim. Acta 2004, 92, 133−138.

(37) https://www.molinstincts.com/chemical-structure/1-1-4-7-7Pentakis-2-pyridylmethyldiethylenetriamine-cstr-CT1103506221. html. (38) Jin, H.-Y.; Akiba, M.; Umakoshi, K.; Sasaki, Y.; Kabuto, K. Dichloro[N,N,N’,N’-tetrakis(2-pyridylmethyl)ethylenediamine]samarium(III) Perchlorate. Cryst. Struct. Commun. 1997, 53, 60−63. (39) Manna, D.; Ghanty, T. K. Complexation behavior of trivalent actinides and lanthanides with 1,10-phenanthroline-2,9-dicarboxylic acid based ligands: insight from density functional theory. Phys. Chem. Chem. Phys. 2012, 14, 11060−11069. (40) Atsushi, W.; Mayayuki, W.; Yoshinori, Y. Synthesis, Crystal Structures and Photophysical Properties of Lanthanide Complexes with Novel Octadentate Oligopyridine Ligands. Rare Earths 2005, 46, 96−97. (41) Marcus, Y. A simple empirical model describing the thermodynamics of hydration of ions of widely varying charges, sizes, and shapes. Biophys. Chem. 1994, 51, 111−127. (42) R, E. N.; Choppin, G. R. Hydration and hydrolysis of lanthanides. In Handbook on the Physics and Chemistry of Rare Earths; Gschneider, K. A., Jr., Eyring, L., Eds.; Elsevier Science Publishers: Amsterdam, 1991; Vol. 15, p 393. (43) Zhang, J.; Heinz, N.; Dolg, M. Understanding Lanthanoid(III) Hydration Structure and Kinetics by Insights from Energies and Wave functions. Inorg. Chem. 2014, 53, 7700−7708. (44) Kuta, J.; Clark, A. E. Trends in Aqueous Hydration Across the 4f Period Assessed by Reliable Computational Methods. Inorg. Chem. 2010, 49, 7808−7817. (45) Heinz, N.; Zhang, J.; Dolg, M. Actinoid(III) Hydration-First Principle Gibbs Energies of Hydration Using High Level Correlation Methods. J. Chem. Theory Comput. 2014, 10, 5593−5598. (46) Wiebke, J.; Moritz, A.; Cao, X.; Dolg, M. Approaching Actinide(+III) Hydration From First Principles. Phys. Chem. Chem. Phys. 2007, 9, 459−465. (47) David, F. H.; Vokhmin, V. Thermodynamic properties of some tri- and tetravalent actinide aquo ions. New J. Chem. 2003, 27, 1627− 1632. (48) Jensen, M. P.; Morss, L. R.; Beitz, J. V.; Ensor, D. D. Aqueous complexation of trivalent lanthanide and actinide cations by N,N,N′,N′-tetrakis(2-pyridylmethyl)ethylenediamine. J. Alloys Compd. 2000, 303-304, 137−141. (49) Ellis, R. J.; Brigham, D. M.; Delmau, L.; Ivanov, A. S.; Williams, N. J.; Vo, M. N.; Reinhart, B.; Moyer, B. A.; Bryantsev, V. S. “Straining” to Separate the Rare Earths: How the Lanthanide Contraction Impacts Chelation by Diglycolamide Ligands. Inorg. Chem. 2017, 56, 1152−1160. (50) Reed, A. E.; Weinhold, F. Natural bond orbital analysis of nearHartree-Fock water dimer. J. Chem. Phys. 1983, 78, 4066−4073. (51) Reed, A. E.; Weinstock, R. B.; Weinhold, F. Natural population analysis. J. Chem. Phys. 1985, 83, 735−746. (52) Ionova, G.; Rabbe, C.; Guillaumont, R.; Ionov, S.; Madic, C.; Krupa, J.-C.; Guillaneux, D. A donor-acceptor model of Ln(III) complexation with terdentate nitrogen planar ligands. New J. Chem. 2002, 26, 234−242. (53) te Velde, G.; Bickelhaupt, F. M.; Baerends, E. J.; Fonseca Guerra, C.; van Gisbergen, S. J. A.; Snijders, G. J.; Ziegler, T. Chemistry with ADF. J. Comput. Chem. 2001, 22, 931. (54) ADF2017, SCM, Theoretical Chemistry; Vrije Universiteit: Amsterdam, The Netherlands, http://www.scm.com. (55) Ali, S. M.; Joshi, J. M.; Singha Deb, A. K.; Boda, A.; Shenoy, K. T.; Ghosh, S. K. Dual mode of extraction for Cs+ and Na+ ions with dicyclohexano-18-crown-6 and bis(2-propyloxy)calix[4]crown-6 in ionic liquids: density functional theoretical investigation. RSC Adv. 2014, 4, 22911−22925. (56) Ali, S. M. Design and screening of suitable ligand/diluents systems for removal of Sr2+ ion from nuclear waste: Density functional theoretical modelling. Comput. Theor. Chem. 2014, 1034, 38−52. (57) Ali, S. M. Thermodynamical Criteria of the Higher Selectivity of Zirconium Oxycations over Hafnium Oxycations towards Organo13115

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116

ACS Omega

Article

the Conductor-Like Screening Model (COSMO) and Its SelfConsistent Generalization to Real Solvents (Direct COSMO-RS). J. Phys. Chem. A 2006, 110, 2235−2245. (79) Xiao, C.-L.; Wu, Q.-Y.; Wang, C.-Z.; Zhao, Y.-L.; Chai, Z.-F. Quantum Chemistry Study of Uranium(VI), Neptunium(V), and Plutonium(IV,VI) Complexes with Preorganized Tetradentate Phenanthrolineamide Ligands. Inorg. Chem. 2014, 53, 10846−10853. (80) Sengupta, A.; Sk, J.; Boda, A.; Ali, S. M. An amide functionalized task specific carbon nanotube for the sorption of tetra and hexa valent actinides: experimental and theoretical insight. RSC Adv. 2016, 6, 39553−39562. (81) Keith, J. M.; Batista, E. R. Theoretical examination of the thermodynamic factors in the selective extraction of Am3+ from Eu3+ by dithiophosphinic acids. Inorg. Chem. 2012, 51, 13−15. (82) Ansari, S. A.; Mohapatra, P. K.; Mohapatra, P. K.; Ali, S. M.; Sengupta, A.; Bhattacharyya, A.; Verboom, W. Understanding the complexation of Eu3+ with three diglycolamide-functionalized calix[4]arenes: spectroscopic and DFT studies. Dalton Trans. 2016, 45, 5425−5429. (83) Leoncini, A.; Ansari, S. A.; Mohapatra, P. K.; Boda, A.; Musharaf Ali, S.; Sengupta, A.; Huskens, J.; Verboom, W. Benzenecentered tripodal diglycolamides: synthesis, metal ion extraction, luminescence spectroscopy, and DFT studies. Dalton Trans. 2017, 46, 1431−1438. (84) Szabo, A.; Ostlund, N. S. Modern Quantum Chemistry: Introduction to Advanced Electronic Structure Theory, 1st ed. (revised); Dover Publications: New York, 1996; pp 1−480. (85) Koopmans, T. Ü ber die Zuordnung von Wellenfunktionen und Eigenwerten zu den einzelnen Elektronen eines Atoms. Physica 1934, 1, 104.

phosphorus Ligands: Density Functional Theoretical Investigation. Eur. J. Inorg. Chem. 2014, 1533. (58) Pahan, S.; Boda, A.; Ali, S. M. Density functional theoretical analysis of structure, bonding, interaction and thermodynamic selectivity of hexavalent uranium (UO22+) and tetravalent plutonium (Pu4+) ion complexes of tetramethyl diglycolamide (TMDGA). Theor. Chem. Acc. 2015, 134, 41−57. (59) Chappa, S.; Singha Deb, A. K.; Ali, S. M.; Debnath, A. K.; Aswal, D. K.; Pandey, A. K. Change in the Affinity of Ethylene Glycol Methacrylate Phosphate Monomer and Its Polymer Anchored on a Graphene Oxide Platform toward Uranium(VI) and Plutonium(IV) Ions. J. Phys. Chem. B 2016, 120, 2942−2950. (60) Sengupta, A.; Bhattacharyya, A.; Verboom, W.; Ali, S. M.; Mohapatra, P. K. Insight into the Complexation of Actinides and Lanthanides with Diglycolamide Derivatives: Experimental and Density Functional Theoretical Studies. J. Phys. Chem. B 2017, 121, 2640−2649. (61) Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648−5652. (62) Lee, C.; Yang, W.; Parr, R. G. Development of the ColleSalvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785−789. (63) Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297. (64) Schäfer, A.; Huber, C.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr. J. Chem. Phys. 1994, 100, 5829−5835. (65) Ahlrichs, R.; Bär, M.; Häser, M.; Horn, H.; Kölmel, C. Electronic structure calculations on workstation computers: The program system turbomole. Chem. Phys. Lett. 1989, 162, 165−169. (66) Treutler, O.; Ahlrichs, R. Efficient molecular numerical integration schemes. J. Chem. Phys. 1995, 102, 346−354. (67) TURBOMOLE V6.6 a development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989−2007; TURBOMOLE GmbH, since 2007; available from http://www.turbomole. com. (68) Dolg, M.; Stoll, H.; Savin, A.; Preuss, H. Energy-adjusted pseudopotentials for the rare earth elements. Theor. Chim. Acta 1989, 75, 173−194. (69) Dolg, M.; Stoll, H.; Preuss, H. A combination of quasirelativistic pseudopotential and ligand field calculations for lanthanoid compounds. Theor. Chim. Acta 1993, 85, 441−450. (70) Dolg, M.; Stoll, H.; Preuss, H. Energy-adjustedabinitiopseudopotentials for the rare earth elements. J. Chem. Phys. 1989, 90, 1730− 1734. (71) Küchle, W.; Dolg, M.; Stoll, H.; Preuss, H. J. Chem. Phys. 1994, 100, 7535. (72) Cao, X.; Dolg, M.; Stoll, H. Valence basis sets for relativistic energy-consistent small-core actinide pseudopotentials. J. Chem. Phys. 2003, 118, 487−496. (73) Cao, X.; Dolg, M. Segmented contraction scheme for smallcore actinide pseudopotential basis sets. J. Mol. Struct.: THEOCHEM 2004, 673, 203−209. (74) Klamt, A. Conductor-like Screening Model for Real Solvents: A New Approach to the Quantitative Calculation of Solvation Phenomena. J. Phys. Chem. 1995, 99, 2224−2235. (75) Klamt, A. The COSMO and COSMO-RS solvation models. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2011, 1, 699−709. (76) Eckert, F.; Klamt, A. Fast solvent screening via quantum chemistry: COSMO-RS approach. AIChE J. 2002, 48, 369−385. (77) Renz, M.; Kess, M.; Diedenhofen, M.; Klamt, A.; Kaupp, M. Reliable Quantum Chemical Prediction of the Localized/Delocalized Character of Organic Mixed-Valence Radical Anions. From Continuum Solvent Models to Direct-COSMO-RS. J. Chem. Theory Comput. 2012, 8, 4189−4203. (78) Sinnecker, S.; Rajendran, A.; Klamt, A.; Diedenhofen, M.; Neese, F. Calculation of Solvent Shifts on Electronicg-Tensors with 13116

DOI: 10.1021/acsomega.8b00933 ACS Omega 2018, 3, 13104−13116