Chem 01201704008
Chem 01201704008
Chem 01201704008
By
Ashutosh Srivastava
CHEM01201704008
DOCTOR OF PHILOSOPHY
Of
HOMI BHABHA NATIONAL INSTITUTE
April, 2023
STATEMENT BY AUTHOR
This dissertation has been submitted in partial fulfillment of requirements for an advanced
degree at Homi Bhabha National Institute (HBNI) and is deposited in the Library to be made
Brief quotations from this dissertation are allowable without special permission, provided that
accurate acknowledgement of source is made. Requests for permission for extended quotation
from or reproduction of this manuscript in whole or in part may be granted by the Competent
Authority of HBNI when in his or her judgment the proposed use of the material is in the
interests of scholarship. In all other instances, however, permission must be obtained from the
author.
Ashutosh Srivastava
ii
DECLARATION
I, hereby declare that the investigation presented in the thesis has been carried out by me. The
work is original and has not been submitted earlier as a whole or in part for a degree /diploma
Ashutosh Srivastava
iii
Dedicated to...........
My Parents
&
Family
iv
Acknowledgement
“Research is an unending quest to unravel the truth, which has widened human vision,
opened newer boulevards, and lightened the dark vague facets of science.”
I would like to thank God, for letting me through all the difficulties. I have experienced your
guidance day by day and without your blessings, this study would not have been possible.
As one flower makes no garland, this dissertation would not have been accomplished without
the support and encouragement of numerous peoples including my family, friends,
colleagues, and well-wishers who contributed in many ways for its successful completion.
First and foremost, I would like to express my deep and sincere gratitude to my guide Dr.
Neetika Rawat for her continuous support, patience, motivation, enthusiasm, erudite
suggestions, and friendly nature throughout the tenure of my Ph.D. I have been fortunate to
have a mentor like her who gave me freedom to explore on my own and to investigate novel
research, and at the same time providing valuable guidance and teaching me how to question
my thoughts and express ideas. A special thanks to Prof. B. S. Tomar for his constant
encouragements and supports; I had started the present thesis work under his direction. I
must thank to Dr. A. K. Satpati for his motivation, valuable advice, discussions and clearing
my scientific doubts.
I would also like to thank my doctoral committee members Dr. S. Kannan, Dr. P. K.
Mohapatra, Dr. T. K. Ghanty and Dr. A. Bhattacharya for their valuable time,
encouragement, critical evaluation of works and insightful comments and suggestions during
the progress review and pre-synopsis presentations. Once again, my sincere thanks to Dr. P.
K. Mohapatra, Associate director RC & I group and Head Radiochemistry division, BARC,
Mumbai for his interest, unlimited support, and motivation. His valuable suggestions for
shaping layout of the chapters of present thesis are highly appreciated.
My sincere thanks also go to my senior colleagues Dr. M. S. Murali, Dr. S. Chaudhury, Dr.
Ankita Rao, Dr. Rama Mohan Rao Dumpala, Dr. Pranaw Kumar and Dr. S. A. Ansari for
their motivation, support, fruitful discussions, unconditional guidance and help whenever I
needed. In addition, I specially want to thank all co-authors in my all publications for their
great scientific contributions. I also thank my lab-mates Shiny S. Kumar, Jisha Pillai, Ritu
Singh and Pooja for their cooperation, moral support and help during this work. I must
thank Dr. Bal Govind Vats and Dr. Praveen Kumar Verma for their support and help.
Finally, words are not sufficient to express my gratitude to my cherished family members who
mean a lot to me. I will always be indebted infinitely to my parents for their spiritually
limitless love, care, sacrifice, motivation, and support throughout my life. I am grateful to my
elder brother (Kinshuk) and sisters (Kanak and Anupama) for their deep love, motivation,
confidence and desired the best for me always. My special heartily thanks to my wife
Anupriya for her unconditional love, endless support, devotion, encouragement, and
continuous faith in me to purse my dreams. As a decent companion she helps me always
overcoming all hurdles of my life. Special thanks to my parents-in-laws for their love and
support. I would like to express my love to my daughter Ashree who gave me strength to
accomplish my Ph.D.
Ashutosh Srivastava
v
Contents
Chapter 1. Introduction 1
1.1 Introduction 2
1.2. Indian Nuclear power Programme 3
1.3. Nuclear Fuel Cycle, types, and Challenges 3
1.4. Radioactive waste, classification, and storage 5
1.4.1. Solid radioactive waste 5
1.4.2. Gaseous radioactive waste 5
1.4.3. Liquid waste 5
1.4.4. Storage of nuclear radioactive waste: Deep geological repository (DGR) 6
1.5. General Properties of Actinides 7
1.5.1. Electronic Configuration 7
1.5.2. Oxidation States, ion types, and redox potential 7
1.5.3. Complexation, hydrolysis, and polymerization 9
1.5.4. Aqueous Speciation 11
1.5.5. Molecular-orbital diagram 12
1.6. Speciation definition and its roles in actinide chemistry 13
1.6.1. Environmental aspects of Actinide chemistry 16
1.6.2. Dissolution and analysis 23
1.6.3. Stability of unusual oxidation state in solution 32
1.6.4. Separation schemes advancement (oxidation states or designing novel 35
ligands)
1.7. Choice of Complexing medium (chelating ligands and novel solvents) 36
1.8. Green and Sustainable Chemistry 43
1.9. Motivation of the work 44
1.10. Scope of the Thesis 46
1.11. Chapters’ distribution 49
vi
2.8. Detection limit of various technique and advantages of 86
electrochemistry for speciation
vii
Chapter 5. Chemical and Redox Speciation of Uranyl ions with three 143
heterobifunctional Phosphonocarboxylate Chelates
5.1. Introduction 144
5.2. Experimental, Instrumentations and Equations 146
5.3. Results and Discussion 146
5.3.1. Protonation of PCs 146
5.3.2. Aqueous speciation (molecular and redox) of uranyl(VI) with PCs 146
5.3.2.1. Molecular speciation (Stoichiometry and Thermodynamic) and physical 146
evidence by UV–Visible spectrophotometry, Calorimetry and ESI-MS
5.3.2.2. Redox Speciation 148
5.3.3. Structural studies-binding mode: (EXAFS and DFT) 152
5.3.3.1. EXAFS study 152
5.3.3.2. Theoretical Study (DFT calculations) 153
5.3.4. Pathways of electron transfer (reduction-mechanism) 156
5.3.5. Aqueous speciation (stability and structure) of uranyl(V) with PCs 156
and Spectroelectrochemical evidence
5.3.6. Aqueous speciation modelling of uranyl(VI) and uranyl(V) with PCs 158
5.4. Conclusions 160
viii
6.3.8.3. Validation of Calibration plots by redox-titrimetric biamperometry 202
6.3.8.4. Scan rate variation and determination of transport and kinetic 202
parameters
6.3.8.5. DES vs. Aqueous 204
6.3.8.6. Reduction-Mechanism 205
6.3.9. In situ spectroelectrochemical Studies 206
6.3.10. Simulation studies 208
Part (c) 209
6.3.11. Redox speciation of Americium in CDHC-MA DES 209
6.4. Conclusions 211
Appendix 244-264
References 265-278
Publications 279
ix
Figures, Tables and Schemes
x
Figure 3.1 Cyclic voltammetry curves: (a) [UO22+]= 10-5 M, (i) in absence of Citrate at 112
pH 4 (ii) in presence of [Citrate]= 5x10-3 M at pH 4 (iii) same as (ii) at pH
6.5; (b) [UO22+]= 10-3 M, (i) in absence of Citrate at pH 4 (ii) in presence of
[Citrate]= 5x10-2 M at pH 4 (iii) same as (ii) at pH 6.5 (c) Square wave
Voltammetry curves with varying pH, Condition: [UO22+]= 10-3 M,
[Citrate]= 1x10-3 M (d) Variation of cyclic voltammograms with scan rates
Figure 3.2 Eh-pH plots, Condition: (a) [UO22+]= 10-5 M, [Citrate]= 5x10-3 M and I= 0.1 113
M NaClO4 (b)[ UO22+] = 10-3 M , [Citrate]= 5x10-2 M and I= 0.1 M NaClO4;
(c) for Peak II (d) for Peak III, both (c) & (d) are at [UO22+]= 10-3 M,
[Citrate]= 1x10-3 M and 0.1 M NaClO4.
Figure 3.3 (a) E vs. logν plot for peak I and (b) i vs. √ν plot (c) is the variation of peak 114
current functions (i/√ν) with scan rates at pH 4 while variation of
chronopotentiometric constant (iτ1/2 ) with i for several values of pH is in
(d), Conditions: (a) to (d) [UO22+]= 10-5 M, [Citrate]= 5x10-3 M and I= 0.1
M NaClO4.
Figure 3.4 (a) E(V) vs. log [1-(t/τ)1/2] plots at pH 4 (b) Variation of 115
chronopotentiometric constant (iτ1/2) with pH at fixed i = 10x10-8 A. (c)
Variation of Peak Current function (i/√ν) with pH at 50 mV/sec scan rate
(d) Variation of E(t=0) with log i at pH 5
Figure 3.5 Plot of i(A) vs. t(sec) with varying pH is in (a) and (b) is the plot of i(A) vs. 116
t-1/2. Condition: (a) and (b) [UO22+]= 1x10-5 M, [Citrate]= 5x10-3 M and I=
0.1 M NaClO4.
Figure 3.6 (a) & (b) are the absorption spectra with varying potential from -0.01 V to - 116
1 V after every 25 min. (c) and (d) are the absorption spectra measured with
varying both potential (from -0.1 V to -1 V) and time of pentavalent
uranium. The appearance of absorption bands are at 732, 743, 756, 774,
785, 798 and 982.8 nm respectively. Condition: (a) to (d) [UO22+]= 4x10-3
M, [Citrate]= 4x10-3 M and I= 0.1 M NaClO4
Figure 3.7 . The Variation of Ep (a) with [UO22+] concentration at fixed [Citrate]= 117
1x10-2 M, I= 0.1 M NaClO4 and pH 4; (b) The variation of F0[UO22+ Conc.]
with [UO22+ Conc.] at pH 4
Figure 3.8 (i) and (ii) are the CV plots (a) UO22+ at 10-5M in absence of PPA and (b) in 119
presence of 5x10-3M PPA conc. (i) at pH 4 (ii) at pH 5. The 3D plots of (iii)
CV at pH 4 and 4.5 (iv) DPV at pH 4 and pH 4.5. The 3D plots of (v) CV
and (vi) DPV at pH 4,7 and 9
Figure 3.9 The Eh-pH plot (i) for reduction peak I and (ii) reduction peak II. (iii) Plot 120
of F0 (PPA) vs. [PPA Conc.]
Figure 3.10 The i(A) vs E(V) plot with increasing scan rate from 50 to 100mV/s (a) at 121
pH 3 ; (b) at pH 5. The E vs logν plot for peak I (c) at pH 3 and (d) at pH 5.
The i(A) vs √ν plot (e) at pH 3; (f) at pH 5.
Figure 4.1 (a) Cyclic voltammetry curves: [Np]=10-5 M and I= 0.1 M NaClO4, Black 136
voltammogram is in absence of phenylphosphonic acid at pH 2.5 while Red
voltammogram is in presence of [PPA]= 1x10-3 M at same pH. (b)
Differential pulse voltammetry curves: at same solution conditions as
described in above.
Figure 4.2 Eh-pH plots, (a) For reduction peak I and (b) For reduction peak II; 136
Condition: [Np]=10-5 M, [PPA]= 1x10-3 M and I= 0.1 M NaClO4.
Figure 4.3 (a) Variation of cyclic voltammograms with scan rates from 20 mV/sec to 137
90mV/sec. (b) E vs. logν plot for Np(V)-Phenylphosphonate (Reduction
peak II); Condition: at [Np]=10-5 M, [PPA]= 1x10-3 M and I= 0.1 M
NaClO4 (c) and (d) are i vs. √ν plot for Np(VI)-Phenylphosphonate
(reduction peak I) and Np(V)-Phenylphosphonate (reduction peak II);
xi
Condition: (a)-(c) at [Np]=10-5 M, [PPA]= 1x10-3 M and I= 0.1 M NaClO4
Figure 4.4 Schematic representation for the complex formation of Np(H2O)5 by 138
PhPO3H2
Figure 4.5 The absorbance spectra (left) and the molar absorbance (right) of the NpO2+ 138
- PPA complexation. ([NpO2+] = 4.23 X 10-5 M with 0.5 M PPA at pH 2)
Figure 4.6 Correlation between the stability constant (log β) and the magnitude of 139
wavelength shift (∆λmax) where, ∆λmax = λ(NpO2L)-λ(NpO2+)
Figure 4.7 Optimised structure of Np(V) and Np(VI) complexes with PPAH- and PPA2- 139
(A) Np(V) with PPAH-- NpO2PPAH (B) Np(V) with PPA2- -NpO2PPA- (C)
Np(VI) with PPAH-- NpO2PPAH+ (D) Np(VI) with PPA2- - NpO2PPA (Red
– Oxygen, Blue – Neptunium, yellow – Phosphorous, Gray – Carbon, white
– Hydrogen)
Figure 4.8 The variation of F0[PPA Conc.] with [PPA Conc.] 140
Figure 5.1 The complexation of uranyl(VI) with Phosphonopropionate (3-PPA) (i) 162
Absorption spectra with increasing the addition of 3-PPA; (ii) The
absorption spectra of individual species: free uranyl(VI) and 1:1 UO22+-3-
PPA complex. (iii) Hyperquad fitted speciation diagram for the
complexation of uranyl with phosphonoacetate (PAA) ([UO22+] = 0.003 M,
[PAA] = 0.02 M).
Figure 5.2 Cyclic Voltammetry Plots of Uranyl(VI) and its complexes with PCs. (i) 163
[UO22+]= 10-5 M (ii-iv) [UO22+] with [PFA], [PAA] and [3-PPA] at 5x10-4
M respectively (v) [PAA]= 5x10-4 M; all at I= 0.1 M NaClO4 and pH=4.
Figure 5.3 (i) Eh-pH plots of [UO22+]= 10-5 M with [3-PPA]= 5x10-4 M, (ii) i(A) vs. 164
E(V) plot with varying scan rates 10-100 mV/sec and (iii) E vs. log ν plot of
UO22+ with 3-PPA, (iv) i(A) vs. √ν plot of UO22+-3-PPA.
Figure 5.4 The different electrochemical plots of UO22+-PAA. (i) is the variations of 165
peak current functions (i/√ν) with scan rates (CV). (ii) is the variation of
chronopotentiometric constants (iτ1/2) with i(A) (CP). (iii) is the plots of
i(A) vs. t-1/2 (CA). (iv) is the E(V) vs. log [1-(t/τ)1/2] plots (CP). i(A) and
E(V) are considered as icom and Ecom.
Figure 5.5 i) Normalized XANES spectra of UO2CO3, U(VI)-PAA, U(VI)-PFA and 166
U(VI)-3-PPA at U L3-edge. (ii) Fourier transformed EXAFS spectra of
U(VI)-PFA, U(VI)-PAA and U(VI)-3-PPA at U L3-edge. The experimental
spectra are represented by Scatter points and the theoretical fit is
represented by Solid line.
Figure 5.6 Optimized DFT structures: (a) Structures for the optimized geometries of 167
Uranyl(VI)- monoprotonated CP complexes. (Colour code of atoms: Blue -
Uranium, Red - Oxygen, Gold - Phosphorous, Grey - Carbon and White -
Hydrogen). (b) Structure for the optimized geometries of uranyl(V)-CP
complex. (Colour code of atoms: Blue-Uranium, Red-Oxygen, Gold-
Phosphorous, Grey-Carbon and White-Hydrogen).
Figure 5.7 Spectroelectrochemical electrolysis data at pH 4. (a) is the absorption 168
spectra with varying cathodic potentials in the wavelength region 375 nm to
550 nm (b) is the absorption spectra with varying cathodic potentials in the
wavelength region 560 nm to 1000 nm
Figure 5.8 Speciation plots of uranyl(VI) and uranyl(V). (a) Uranyl(VI) with 3-PPA 169
(H- is the -OH group, M is uranyl(VI) and L is 3-PPA) (b) uranyl(V) with
CP (PAA), citrate and carbonate chelators (c) percentage of pH region of
uranyl(v) stability with organic chelators.
xii
Figure 6.1 Characterization: (a) IR spectra of UO3 and UN dissolved Maline solution. 214
TGA plots (b) Wt% loss versus Temperature (c) The UV-Visible absorption
spectra of UO3 dissolved in Maline solution, blank Maline and UO22+ and
malonic acid complex in aqueous solution.
Figure 6.2 Redox speciation of Uranium matrices in Maline. Cyclic voltammetry plots 215
of (a) UO3, UN, UO2, U3O8 and blank Maline (b) U std., U metal and U3Si2
(c) PuO2 and (U, Pu)O2 (d) Scan rates variation studies by Cyclic
voltammetry (CV) on UO3 dissolved in Maline.
Figure 6.3 Insitu UV-visible Spectroelectrochemical Data. (i) Absorbance spectra of 216
U(VI) in Maline with applied cathodic potentials. (ii) and (iii) are the
appearance of new absorption peaks with applied cathodic potentials related
to U(IV) and U(V) species in Maline respectively.
Figure 6.4 (i) Normalised XANES spectra at U L3-edge along with standard samples. 217
(ii) Fourier transformed EXAFS spectra of U sample at U L3-edge in (a) R
space and (b) K space along with the best fit. The experimental spectra are
represented by Scatter points and the theoretical fit is represented by Solid
line.
Figure 6.5 (a) Equilibrated structure of the Maline system containing Uranyl ion in the 218
presence of water molecule. System representation with 5% water. color
code: Steel: UO2 species; Red & White: H2O; Cyan: Cl; rest are malonic
acid and choline (b) Optimized structures of [UO2-(MA)2-(H2O)2]2+ (c) The
calculated energy levels (eV) of the MOs of the [UO2-(MA)2-(H2O)2]2+
complex using B3LYP/TZ2P/ZORA level of theory.
Figure 6.6 (a) The luminescence intensity of Uranyl with increasing concentration of 219
Maline (b) Comparison of luminescence intensity of Uranyl with Malonic
acid and Maline at two different concentrations (c) The absorption spectra
for Uranyl titration with Maline (d) Deconvoluted spectra representing the
absorption profile for the pure components (e) Raw calorimetric data for the
interaction of Uranyl with Maline (f) Integrated Data showing the
cumulative heat released on successive additions of DES solution to fixed
volume of Uranyl solution; [UO22+] = 0.01 M, [DES] = 0.48 M
Figure 6.7 (a) Differential pulse voltammetry plots with increasing concentration of 220
UO3 (b) Calibration Plot of UO3 dissolved in Maline solution obtained with
increasing concentration (c) Electrochemical interference study on UO3
dissolved Maline with other metal ions dissolved Maline (d) Selectivity of
Maline for different dissolved metal ions.
Figure 6.8 (a) measured density of DES 1 with varying temperature and fitting (b) 225
measured viscosity of DES 1 and fitting (c) Thermograms obtained in DSC
for pure components.
Figure 6.9 (a) IR spectra of the UO3 and UO2 dissolved in DES 1 (b) UV-Vis spectra 226
of UO3 dissolved DES 1 (c) TGA data of the UO3 and UO2 dissolved in
DES in terms of the wt % loss with temperature (d) derivative plot (e)
electrochemical potential window and redox peaks of DESs.
Figure 6.10 (a) and (b) are the CV and DPV plots for the oxides of uranium dissolved in 227
DES 1. (c) and (d) are the CV and DPV plots for the oxides of uranium
dissolved in DES 2. The concentration of uranium oxides ~ 4.5 to 5 mg/ml.
Figure 6.11 (a) and (b) are the DPV plots with increasing concentration of UO3 (peak I 228
and II) in DES 1. (c) and (d) are the ip(A) versus conc. UO3 plots in DES 1
(e) Ep vs. concentration of UO3 in DES 1
Figure 6.12 (a) is the recorded DPV plots for U3O8 by varying its concentration (DES 229
2). (b) (i) to (iii) are the ip(A) vs conc. U3O8 plots for reductive peaks I to
III.
xiii
Figure 6.13 (i) and (ii) are the scan rates variation studies on uranium oxides dissolved 230
in DES 1 in cyclic voltammetry (CV). (a) is the CV plots recorded with
varying scan rates. (b) is the E vs. logν plots. (c) is the ln|ip| vs. Ep-Ef plots
(d) is the i(A) vs. √ν plots of their respective uranium oxides.
Figure 6.14 In situ UV-visible Spectroelectrochemical Data. (a) Absorbance spectra of 231
U(VI) in DES 1 with applied cathodic potentials. (b) and (c) are
appearances of new absorption peaks with applied cathodic potentials
related to U(V) and U(IV) species in DES 1 respectively.
Figure 6.15 In situ UV-visible Spectroelectrochemical Data. (a) Absorbance spectra of 232
U(VI) in DES 2 with applied cathodic potentials with the isosbestic point.
(b) (i) to (iii) showing the appearance of new absorption peaks with applied
cathodic potentials related to U(V) species in DES 2 respectively
Figure 6.16 MD simulation results: (a) Geometries of (i) C8mimBr and (ii) malonic acid 233
molecules. (b) Enlarged picture of DES-uranyl system. (i) UO2-Br2, UO2-
Br3 and UO2-Br4 and (ii) UO2-2malonic acid. Here, U: pink; O: Red; Br:
light blue; malonic acid: Blue. DFT simulation results: (c) Optimized
structures of [UO2-(MA)2-(Br)2] (d) The calculated energy levels (eV) of
the MOs of the [UO2-(MA)2-(Br)2] complex using B3LYP/TZ2P/ZORA
level of theory.
Figure 6.17 (i) (a) Latimer diagram for Am in 1 M perchloric acid (b) dissolved sample 234
of Am-oxide in DES (c) UV-Visible absorption spectra of Am in DES (ii)
(a) CV plots of Am-oxide in CDHC-MA and blank CDHC-MA (b) DPV
plots with increasing concentration of Am in CDHC-MA (c) CV plots with
varying scan rates from 10 mV/sec to 150 mV/sec (d) the ip vs. √ν plot
xiv
Table 4.5 Bond distances (Å) in all the Np species involved in complexation process 142
Table 5.1 Protonation constants for three Phosphonocarboxylates. (Experimental 170
conditions: T = 298 K, P = 0.1 M Pa and I =1.0 M NaClO4).
Table 5.2 Assignment of Species from m/z values obtained in ESI-MS spectra. 171
Table 5.3 Thermodynamic parameters for the complexation of Uranyl with three 172
PCs. (Experimental conditions: T = 298 K, P = 0.1 M Pa and I =1.0 M
NaClO4, [UO22+] = 0.003 M, [L] = 0.02 M; L = PFA/PAA/3-PPA).
Table 5.4 Summary of Cyclic Voltammetry data of UO22+-PCs and comparison of 172
potential of U(VI)/U(V) redox couple with various complexing chelators.
Table 5.5 Summary of electrochemical parameters obtained from fitting of Plots. 173
Table 5.6 Bond length, coordination number and disorder factor obtain by EXAFS 174
fitting for U(VI)-PCs at U L3-edge.
Table 5.7 The binding energy (B.E) and free energy of formation for all the three 175
uranyl(VI)-CP complexes in gaseous as well as aqueous phase
Table 5.8 The calculate bond distances (in A0) between the uranium and the 175
coordinating atoms in stable optimized geometries of uranyl(VI)-CP
complexes.
Table 5.9 Stability constant, bond distances and partial charges of Uranyl(V)-CP 176
complexes
Table 6.1 Electrochemical thermodynamic and kinetic parameters of U matrices in 221
Maline. (Irrev. = Irreversible redox behaviour)
Table 6.2 Bond length (R (Å), phase-corrected), coordination number (N) and 222
disorder factor (s2) obtain by EXAFS fitting.
Table 6.3 Analytical figures of merit and validation of developed analytical 222
methodology for Uranium matrices sample determination in Maline.
Table 6.4 Comparison of Green metrics of present methodology (direct sequestering 223
and analysis of uranium matrices with Maline supramolecular scaffold)
and previous method (i.e., hazardous acidic/additives dissolution and
biamperometery analysis).
Table 6.5 The observed and calculated weight loss along with the liberated 235
molecules at various temperatures in TG-DSC.
Table 6.6 Redox thermodynamic, Diffusion coefficient and kinetic parameters of 236
UO3 and UO2 in DES
xv
Scheme 5.2. Coordination modes of PC chelators 161
complexes
xvi
List of Abbreviation
xvii
SYNOPSIS
Nuclear energy, a mature technology and low carbon energy source, is going to be one of the
major sources of energy as the demand for energy and concern of reducing harmful
greenhouse gas emission grows worldwide. Actinides viz., uranium (U) and plutonium (Pu)
are used as a fissile material for nuclear reactor, which on irradiation in nuclear reactor
generates the large volume of spent nuclear fuel (SNF) consisting of highly radioactive
unused fuels, minor actinides, and fission products. The reprocessing and safe management
of nuclear waste of SNF are important concern for sustainable development of nuclear
energy. Aqueous (Aq) solution occupies a strategic place in both front and back end of
nuclear fuel cycle (NFC) related to (1) hydrometallurgical processing and reprocessing
namely, dissolution, solvent extraction, and recovery of actinides, and (2) assess the safety of
nuclear waste disposal. Non-aqueous (Non-Aq) solvents such as ionic liquids and deep
eutectic solvents etc. are the new generation of media being explored globally for various
industrial applications including nuclear industry. The solution chemistry of the early actinide
elements (U, Np, Pu and Am) are of vital importance and interesting to the nuclear industry.
One of the important feature of early actinide elements is the formation of actinyl ions
(AnO2n+; n= 1, 2). Actinyl ions are the important species in the nuclear fuel cycle because of
their multifaceted and highly characteristic chemical behavior [1-4]. These ions have dynamics
behaviour in Aq solution and can form stable complexes and nanoclusters with various
organic and inorganic ligands in Aq and Non-Aq solvents. The chemical behaviour of actinyl
hydrolysis, solubility product, coordination modes, redox potentials, and stability constants [1-
4]
. Thus, understanding the solution chemistry of early actinides, particularly actinyl ions, in
Aq and Non-Aq are crucial. Novel research with the aim of probing direct metrical
xviii
information about redox and molecular speciation of actinyl ions in both Aq and non-Aq as a
function of physicochemical conditions are quintessential requirement for (1) describing in-
depth solution chemistry, (2) efficient design of separation processes, and (3) fundamental
understanding migration of actinyl ions in environment. This will help in the advancement of
reprocessing of SNF and the geological disposal strategies of nuclear wastes and to mitigate
challenges of NFC.
and pyrochemical processes. The complexity of these processess and resulting technical
challenges toward sustainable nuclear energy necessitates various research activities related
to the development of alternative methods based on Non-Aq, viz. ionic liquids (ILs) and deep
eutectic solvents (DESs). Furthermore, for chemical quality control, nuclear material
accounting, and nuclear forensic applications, the routine analysis of nuclear materials (NMs)
development of novel fast dissolution methods using less hazardous chemicals to avoid time
generation of harmful NOx gases. Non-Aq method for reprocessing of SNF and dissolution of
NMs has numerous advantages, such as minimum waste generation, excellent thermal and
radiation stability, lower criticality risk, compact infrastructure, and wide electrochemical
windows. DESs are the nascent entries to the repertoire of Non-Aq chemical media and
organic solvents and ionic liquids in the 21st century. DESs have emerged as a new type of
designer solvents and have synthesized by combining a hydrogen bond acceptor (HBA) and a
hydrogen bond donor (HBD), which can form eutectic mixture with large depression in
[5]
melting point through extensive hydrogen bonding interactions . DESs have many
xix
favorable characteristics such as easy preparation, a large electrochemical window
physicochemical properties. DESs can be used to diverse filed of research viz., metal and
actinides. IUPAC defines speciation as “species distribution based on their oxidation states
solutions depends basically on three main factors (1) Actinide redox processes, (2) Solid
actinide phases, and (3) Complexation reactions [1-4]. The actinide speciation in solution is a
daunting task and are highly challenging because (1) unlike most elements of the periodic
table, there is absence of geologic precedent, (2) their diverse chemistries, including rich
complexes formed with the actinide cations in these states vary for Aq and Non-Aq solutions.
In addition, speciation of actinides is very mobile and the effect of pH on their speciation is
still couched. Based on aforesaid reasons, the independent knowledge of chemical behavior
Therefore, attempts should be made to provide cohesive database on redox and molecular
actinides. Speciation is a cornerstone of different facets of nuclear fuel cycle and it may be
Dissolution, analysis, and recovery in non-aqueous medium (ILs and DESs), (iii) Stability of
unusual oxidation state in solution, and (iv) Separation schemes advancement (oxidation
xx
(i) Actinide environmental aspects (aqueous chemistry): The presence of chelating ligands
their redox, hydrolysis and diffusion behavior and thereby retards or enhances the migration.
The conversion
ersion from an oxidized/potentially mobile form to a more reduced/potentially less
dissolution in ILs or DESs since it affects the metal matrices processing and electrochemical
behavior, such as solubility, dissolution mechanism, complexation and interaction with their
constituents,, and redox potentials [5]. Metal species dissolved in DESs will experience a very
strong ionic atmosphere and hence its chemistry fundamentally will be different from
traditional molecular solvents. Foreseeing the applications of DESs as a novel media, probing
pr
the mechanistic redox and molecular speciation are essential to understand the chemistry and
the mechanism that is occurring in processing and solvent extraction and the redox
xxi
(iii) Stability of unusual oxidation state in solution: Pentavalent uranium (UVO2+) has
historically been reported as unusual (rare) oxidation state of uranium in the comprehensive
1969 review of “U(V) chemistry” by Selbin and Ortego et al. [6] in Aq solution. It is unstable
chemistry and a knowledge gap related to the comprehensive understanding of its stability
and structure in the Aq solution. Its Aq speciation plays a crucial role in the solubility and
mobility of uranium in the environment and in investigating the mechanism of biotic and
ligands): Separations of actinides can be achieved either by (a) determination and control of
actinide in different oxidation states or (b) designing novel ligands for complexation of
actinides/actinyl ions. (a) Early actinides (U, Np, Pu and Am) exist in more than one
oxidation states and every oxidation state has different chemical behavior in solution.
Separation techniques/methods i.e., PUREX and SESAME exploit the distinct variation in the
chemical makeup
and behavior of
different oxidation of
actinides in solution
feasible to easily
partition/separate
another. (b) Chelating ligand with stereognostic coordination platform tends to form an
intramolecular hydrogen bond (IMHB) with the oxo group of the actinyl ions and imparts
xxii
[7-8]
specificity and/or selectivity towards actinyl ions binding . Therefore, to develop and
Choice of Complexing medium (chelating ligands and novel solvents): The principles
about choice of chelating ligands, chemical media, and ligand designing should be depending
on their relevant application and important role in mind so that new research activities can
redox control, and actinyl-oxo bond (i.e., O=U=O) activation. Keeping these in mind, in the
present dissertation, the chelating ligands and chemicals media, as discussed below [Fig. 2],
have been prudently chosen to cover all speciation aspects discussed under title (i) to (iv) and
their respective “redox and molecular speciation” data with “actinyl ions” were generated.
systematic study has been carried out on the speciation of UO22+ in citrate medium
considering its redox behavior. In addition, many discrepancies exist among published data
and a knowledge gaps remain related to its accurate speciation prediction in citrate medium.
The redox speciation of U with phosphonate-based ligands is not available in literature. There
species are scarce in literature. The aqueous speciation, solution thermodynamics, redox
phosphonocarboxylate ligands have not been reported in literature. Literature is scarce on the
electrochemical studies on f-block elements in DES. Literature survey further implies the
lack of direct sequestration of U ions from their diverse solid matrices and their fundamental
chemical aspects (i.e., speciation) in Maline, imidazolium cation based DESs. Furthermore,
the green analysis methodology for U in DES has not been reported.
xxiii
Chapter 1: Introduction: This chapter start with providing some background
background about the
growing demand for nuclear energy, nuclear programme in India, nuclear fuel cycle (NFC),
nuclear waste and its types, and the various pathways of actinides release into the
definition of speciation is presented. In addition, the redox and molecular speciation was
described in details. The scenario of possible interaction of actinides, which governs the fate
and transport of actinides in environment, are discussed. A brief discussion about important
role of speciation of actinides in various area of NFC is discussed in terms of four major
applications. The role of complexing agents towards actinides speciation is highlighted. The
explanation for choice of different complexing agents is given. Finally, the motivation for the
Chapter 2: Experimental and Instrumentations: In this chapter, the details about the
chelators, deep eutectic solvents, used in the present thesis are discussed. The source of the
performed by multi-
A brief description about each technique used for speciation studies are discussed. Since
xxiv
to calculate physicochemical parameters, stability constant, energetics, green chemistry
metrics, coordination numbers and bond distances of actinyl ions are discussed.
(carboxylate and phosphonate based organic chelators): In this chapter, the speciation of
uranyl (UO22+) with (1) citrate and (2) phenylphosphonic acid (PPA) were investigated on the
basis of its redox behaviour using electrochemical methodology with varying physiochemical
conditions. The UO22+ in citrate media exists as UO22+, monomer [UO2Cit]-, dimer
[(UO2)2Cit2]2- and polymeric species. [UO2Cit]- and [(UO2)2Cit2]2- were found to be reduced
in two irreversible one-electron steps, with a chemical reaction coupled between them (ICI
reduction mechanism), to a stable [UIVCit]+ species at pH ≤ 5 and via a single reduction step
techniques were applied to corroborate the finding of an ICI reduction mechanism and to
evaluate electron-transfer kinetic (k0 and α) parameters. In the presence of PPA, the UO22+
has been found to form 1:1 dominant species, i.e., UO2PhPO3, over a wide pH range. Cyclic
voltammetry and differential pulse voltammetry were used to investigate the redox behavior
of UO2PhPO3 species, Eh-pH plot and electron-transfer kinetic (D0, ks and α) parameters.
Using the Deford–Hume formula, the stability constant (log β) of [UO2Cit]- and UO2PhPO3
species were calculated to be (6.8 ± 0.11) and (6.98 ± 0.12), respectively. ESI-MS studies on
solutions of varying UO22+: citrate/PPA ratios and pH were used to confirm complex
chemical reaction interposed between the two-electron transfer steps and to provide evidence
of neptunium with Phenylphosphonic acid: This chapter aimed at investigating the redox
and complexation of Np with phenylphosphonic acid (PPA) using electrochemical and UV–
xxv
Visible absorption spectroscopy and DFT calculations. The cyclic voltammetric (CV)
measurement indicates the complexation of both oxidation states (V & VI) of Np with
[PhPO3]2-. The presence and stability region (via Eh-pH plot) of new species of Np-PhPO3 in
two different oxidation states (V & VI) in aqueous solution are identified. The kinetics for
electron-transfer kinetic (D0, ko and α) parameters by CV. The spectral parameters, namely
λmax and its molar extinction coefficient (ε) are obtained from UV-Visible spectrophotometric
showed the formation of 1:1 complex only and the stability constant (logβ) obtained for the
same from both the methods found to agree with each other. DFT calculations are carried out
the most probable coordination mode and energetic of complex formation reaction.
Chapter 5: Chemical and Redox Speciation of Uranyl ions with three heterobifunctional
with three phosphonocarboxylate ligands, which are differing in CH2 spacer, were probed
ESI-MS studies demonstrate the physical evidence of the dominance of MLH type of
complexes. Formal (Efcom) and peak (Epcom) potentials, Pourbaix diagrams (Eh-pH plots),
heterogeneous electron-transfer kinetic (D0, kof and αn) parameters and mechanistic electron
transfer process of complexes were acquired from cyclic voltammetry (CV). Structural
bond between -OH of ligands and axial oxygen of UO22+. Interestingly, the absence of
xxvi
disproportionation of uranyl(V) and hence its aqueous redox stability with
pentavalent uranyl(V) species with the phosphonocarboxylates in the aqueous was proposed
and their stability and structure were determined. Spectroelectrochemical measurements were
employed to provide evidence of existence of uranyl(V) species and its electronic absorption
spectra.
dissolution, speciation, and analysis of uranium (U) and americium (Am) oxide were probed
in deep eutectic solvents (DESs). Seven U matrices (UO3, UN, UO2, Rb2U(SO4)3, U metal,
U3Si2, and (U, Pu)O2 were dissolved in Maline without any external additives. New DESs
acid (MA) and diglycolic acid (DGA) as HBD, and (2) Choline dihydrogen citrate and
malonic acid (CDHC-MA) in 1:1 ratio. Dissolution of uranium oxides (UO3 and UO2) were
explored in imidazolium based two DESs. The cyclic voltammetry, differential pulse
CDHC-MA DES was carried out only to know the feasibility of its conversion to higher
oxidation state in the form of actinyl ion. Redox thermodynamics (peak potential and formal
redox potential), transport property (diffusion coefficient D0), electron transfer kinetic
parameters (αn and k0) and mechanistic electron transfer of the dissolved U in above DESs
were investigated. Under molecular speciation, the structural analysis on UO3 dissolved in
Maline was conducted by EXAFS. Theoretical simulations (MD and DFT) were performed to
support the experimental results and to acquire diffusion, optimized structure, probable
equatorial coordinating atoms, binding energy, and molecular orbital diagram of U species in
xxvii
provide information about the interaction of uranyl with Maline. Green analysis methodology
determination of U in nuclear material samples and green chemistry metrics were calculated
Chapter 7: Summary and Future outlooks: This chapter provides the summary about the
results of the studies performed as a part of the thesis and suggest some future scopes relevant
and Np, which are quite persistent over a wide range of Eh-pH. The order of stability constant
(logβ) and redox potential of uranyl with these ligands were found to be as
tendency towards UO22+. The association of couple chemical reaction was discovered with
electron transfer (ET) steps for UO22+ complexes with citrate and phosphonocarboxylate,
respectively. Deep eutectic solvents namely Maline and imidazolium cation based DESs
shows encouraging solvation potential for diverse kinds of U matrices. Furthermore, these
DESs work as designer solvents for quantitative green analysis of U based NMs.
Interestingly, the pentavalent uranyl (UO2+) species was detected to be redox stable in a
respective stability constant (logβ), structure and the electronic absorption spectra were
reported. As future research, development of new methodologies for (1) redox based selective
group of actinyl, (3) isolation of pentavalent uranyl species, and (4) actinides’
xxviii
electrodeposition, electrocatalytic recovery, extraction, and separation using DESs, are
anticipated.
References:
1. Silva, R. J., and H. Nitsche. Radiochimca Acta 70, no. Supplement (1995): 377-396.
2. Altmaier, Marcus, Xavier Gaona, and Thomas Fanghänel. Chemical reviews 113, no. 2
(2013): 901-943.
3. Maher, Kate, John R. Bargar, and Gordon E. Brown Jr. Inorganic chemistry 52, no. 7
(2013): 3510-3532.
4. Jones, Matthew B., and Andrew J. Gaunt. Chemical Reviews 113, no. 2 (2013): 1137-
1198.
5. Smith, Emma L., Andrew P. Abbott, and Karl S. Ryder. Chemical reviews 114, no. 21
(2014): 11060-11082.
6. Selbin, Joel, and J. D. Ortego. Chemical Reviews 69, no. 5 (1969): 657-671.
7. K. N. Raymond et al. 1992. Journal of the American Chemical Society, 114(21), 8138-
8146.
xxix
Chapter 1
Introduction
Aim: This chapter aimed to provide some background about the (i) growing demand of
nuclear energy and nuclear power programme in India, (ii) nuclear fuel cycle (NFC) and
associated challenges, nuclear waste, and its types and (iii) general properties of actinides.
The term “speciation” is defined and the important role of speciation of actinides in various
area of NFC is briefly discussed. The choice of complexing agents presented with their
substantial applications. The terminology of green and sustainable is described. Finally, the
motivation of works, scope of the dissertation and chapters’ distribution are given.
1
1.1. Introduction
energy source, which has the possibility of reducing greenhouse gas concentration in the
atmosphere and capable of meet future energy needs [1-2]. Nuclear energy is a form of energy
released from the nucleus in two ways: fission or fusion. As for 2022, the nuclear energy is
used by ~ 30 countries and amounts for a total of 11% of the worldwide produced electricity
[3-5]
. Because of rise in global electricity demand and to avoid the drastic change in climate
from negative impacts of fossil fuels, the production of electricity from non-fossil fuel source
viz., nuclear, wind, hydro solar etc., is in high demand. The nuclear fuel materials have fissile
235 233 239 [6]
isotopes such as U, U and Pu , which on interaction with neutrons undergo fission
and released enormous amount of energy (200 Mev). In comparison to fossil fuels, nuclear
energy requires less quantities of fuel. From the nuclear energy industry, an estimate is to
[5]
increase the total electricity production for the year 2050 ranges between 24%-184% .
Unlike fossil fuel-based electricity production, the side product of the nuclear power reactor,
the “burnt” nuclear fuel (called as spent nuclear fuel (SNF)) must be reprocessed and/or
contained and maintained for a period long enough to allow the long lived toxic radioactive
nuclides to decay towards their decay chain in final non-radioactive form due to its long-term
and high radiotoxicity [7-8]. These radioactive nuclides are mostly fission products and trans-
uranic elements (minor actinides (MAs) and plutonium) which are produced as a side product
of the reactor nuclear reaction. Thus, benefit of nuclear energy production comes with the
inevitable hazards of long-lived radionuclides and a major drawback of nuclear energy is the
disposal of SNF. Therefore, the reprocessing of SNF and management of nuclear waste
became the key issues for the sustainable development of nuclear energy.
2
1.2. Indian Nuclear power Programme: Dr. Homi Jahangir Bhabha (initiator of Indian
atomic energy program) envisaged a three-stage Indian nuclear power program in 1954, as
Stage 1: Natural uranium is used as fuel uranium and heavy water as moderator and coolant
in pressurized heavy water reactors (PHWR) to generate electricity and 239Pu. The spent fuel
obtained from PHWRs is reprocessed to recover the residual uranium (U) and plutonium (Pu)
Stage 2: This stage is based on the fast reactor technology and involves utilization of
plutonium recovered from PHWRs. Plutonium obtained from the Ist stage is mixed with
uranium (as a mixed-oxide (MOX) or as metallic fuel) and used in the Fast Breeder Reactors
238
(FBRs). For power production, the FBRs will fission Pu and breed more it from the U. In
FBRs, the fertile Thorium (Th) used as a blanket to breed more fuel via producing 233U.
Stage 3: This stage plan to achieve the power generation from vast thorium reserves, which
233
found in India, in place of U. It utilizes the fissile U, generated in the IInd stage, as a fuel.
232
Power will be generated from the Th-233U fuel aided by plutonium and this could be
3
Nuclear fuel cycle (NFC) is a series of industrial process involving various activities
[7]
[Fig. 1.2] to produce electricity from uranium in a nuclear power reactor . The cycle starts
with the milling mining of uranium and ends with the disposal of nuclear waste in a proper
way. Being a strategic material and radioactive in nature, the used nuclear fuel needs to be
taken care of for reuse and disposal. Based on various activities, as shown in Fig. 1.2, the
NFC is mainly classified in two parts ‘front end’ and ‘back end.’ The front end is associated
with the preparation of the fuel and the ‘service period’ in which fuel is used during reactor
[7]
operation . The steps associate with front end is mining and milling, conversion,
enrichment, and fuel fabrication. Nuclear fuel burns in reactor to produce electricity and
generates unused nuclear fuel which is known as SNF. The back-end of NFC basically deals
with the safe management of SNF including its temporary storage, reprocessing, reuse, and
disposal of radioactive waste via various treatment to lower its radioactivity. Based on the
management of spent fuel, NFC further can be categorized into two types viz., open, and
closed fuel cycles. If spent fuel is not reprocessed, the NFC is referred to as ‘open’ or ‘once-
through’ fuel cycle while spent fuel is reprocessed, and recycled, it is referred to as ‘closed’.
4
The major challenges which NFC and nuclear energy production faces are (i) the
requirement for continuous enhancement of safety and safety culture to avoid nuclear
accident, (ii) the need to control the spread of technologies and materials that may be used for
non-peaceful purposes, (iii) the storage of SNF and the appropriate operation of reprocessing
and recycling to recover unused fissile materials, (iv) the increased fission product,
transuranic and activation product inventories from high burn-up along with higher decay
heat and neutron sources, and (v) the proper management and disposal execution of highly
[8]
radioactive wastes, especially for HLW . The separation of minor actinides from HLW is
also difficult task. To make the nuclear sector a substantial contribution to meeting the
world’s energy demand, above challenges require consistent effort to resolve, nuclear fuel
national and international efforts should be to enhance the safeguards and non-proliferation.
arises in a wide range of concentrations and in a variety of physical and chemical forms viz.,
solid, liquid and gases. Based on their form the radioactive wastes are classified as below:
1.4.1. Solid radioactive waste: These wastes are classified as combustible, non-
1.4.2. Gaseous radioactive waste: It includes gaseous fission products generated due to
radiolysis.
Depending on radioactivity, the category of solid and gaseous wastes is given in Table 1.1.
1.4.3. Liquid waste: Based on level of radioactivity and as per IAEA safety and protection
5
Table 1.1. Classifications of Solid and Gaseous waste [11]
Solid Gaseous
Category
Surface Dose (mGy/hr) Activity Level (Bq/M3 )
I <2 < 3.7
II 2-20 3.7- 3.7x 104
III >20 >3.7 x 10 4
IV Alpha bearing
(i) Exempt Waste (EW): Waste that follows the criteria for clearance, exemption, or
(ii) Very Short-Lived Waste (VSLW): This kind of waste primarily is radionuclides with very
short half-lives. Waste that can store for decay for a limited period upto a few years and
(iii) Very Low-Level Waste (VLLW): Waste of this kind does not meet the criteria of EW and
does not require a high level of containment and isolation. It is suitable for disposal in near
(iv) Low Level Waste (LLW): Waste under this type is above clearance levels and requires
robust isolation and containment for periods of up to a few hundred years and is appropriate
(v) Intermediate Level Waste (ILW): Waste that requires a greater degree of containment and
isolation than that provided by near surface disposal and it needs disposal at greater depth.
(vi) High Level Waste (HLW): Waste with (i) large amounts of long-lived radionuclides that
need to be considered in the design of a disposal facility or (ii) levels of activity concentration
high enough to generate significant quantities of heat by the radioactive decay process.
1.4.4. Storage of nuclear radioactive waste: Deep geological repository (DGR): Deep is
the preferred route for the final disposal of the nuclear wastes containing SNF and HLW. The
6
nuclear wastes are placed inside DGR which is based on a multibarrier system containing
engineered barrier system and natural geological barrier for complete isolation of radioactive
waste from biosphere for hundreds of thousands of years. The host rock constitutes the
natural geological barrier whereas the engineered barrier system consists of waste form,
canisters, overpack, buffer and backfill material. For long-term safety of DGR, natural
geological barrier should have various features such as high strength, high thermal
conductivity, low thermal expansion, high sorption capacity, no ingress of water from
surrounding and minimum interaction with canister and fluid while the backfill/buffer
material should have high swelling potential, high radionuclide retention capacity, low
[12]
hydraulic conductivity, and low gas permeability . Though the release of actinides from
the repository to biosphere is negligible however due to decay heat and corrosion of canisters,
the release of radioactive actinides to the biosphere through aqueous medium is anticipated.
1.5.1. Electronic Configuration: The actinides correspond to the filling of the 5f electron
Actinides (Ac to Np) show greater tendency to retain 6d electrons due to smaller energy
lanthanides. 5f orbitals of Pu and subsequent actinides are of lower energy than 6d orbitals,
1.5.2. Oxidation States, ion types, and redox potential: In contrast to lanthanides, the
actinides, particularly early elements of the series exhibit more variation in their oxidation
states and have multiple oxidation states such as +4, +5, +6 and +7. This is due to the close
proximity of the energy level of the 7s, 6d, and 5f electrons. In aqueous solution, the
ionization energy and hydration energy are two important factor that are responsible for
the conversion of one oxidation state to another and the most stable oxidation state of
7
actinide is decided by a combination of these two factors. Actinides in second half of the
series, i.e., beyond americium, shows +3 as the most stable oxidation state because with
increasing atomic number the 5f-orbital is comparatively stabilized than the 6d and 7s.
Hence, the second half of actinides behaves like lanthanides. The prevalent oxidation
states of the actinides are given in Fig. 1.3. The actinides in +3 and +4 oxidation states
generally exists a M3+ and M4+ hydrated ions in solution whereas in higher oxidation
states such as +5 and +6, due to very high ionic potential, the actinide ions form the
oxygenated species of the general type MO2+ and MO22+, which are named as actinyl
ions. The actinyl ions are remarkably stable due to strongly covalent bonded symmetrical
linear (or very nearly linear) O=M=O ions and during variety of chemical transformation
it persists as a single unite. The formation of oxygenated species decreases the effective
charge on actinyl ions and thus the order of effective charge on different oxidation states
oxidation mostly found in case of Np and Pu in alkaline aqueous solution in the form of
MO4(OH)23-. The redox potentials of early actinides are shown in Fig. 1.3. The redox
NpO2+, and Pu4+, and (ii) small difference in redox potential, especially in case of Pu,
disproportionation reaction of pentavalent actinyl ion (MO2+) ion, where M= U and Np,
constant (log K) value at 25ºC for disproportionation of UO2+ and NpO2+ are reported to
be 9.30 and -6.72, respectively [15]. The presence of hydrogen ions (i.e., high acidity) and
complexing ions (such as F- and SO42-), which can complex strongly with M4+ and
8
MO22+ ions have dominant effect on disproportionation reaction of MO2+ ion. The
Pourbiax diagram, (Eh-pH plot), is a useful tool to know the dominancy of actinides’
for actinide. Water itself has oxidation and reduction characteristics which provides boundary
on the oxidation states stability of actinide in aqueous solution. Thus, the first step for
constructing the Pourbiax diagram for actinide is to get an idea about the chemical stability
area of water (oxidation and reduction). The upper limit of water is defined by the potential
when the oxygen generation starts on the anode according to reaction ½ O2 + 2H+ +2e- =
H2O. The relation between potential and pH is E = 1.299 – 0.0591 pH. The lower stability
limit of water is defined by potential where hydrogen gas (H+ + e- = ½ H2) is evolved and at this
[16]
condition the corelation between potential and pH is E = 0.059 pH . The actual overall
boundaries of the Pourbiax diagram is described by the above stability field of water.
hard Lewis nature, the actinide-ligands bonds are mainly ionic, kinetically labile and non-
9
directional and show preferences to binds with hard Lewis bases such as fluoride and
oxygen. The presence of two oxo group in the linear in the actinyl ions provide a steric limitation
on the number of donor groups for binding with these ions. The oxo group oriented in axial
position and coming ligands will binds through equatorial position in actinyl ions. In absence of
attraction, i.e., ionic potential (Z/r; Z= effective charge on the ion and r= ion size) and its order
This order is so because the -yl oxygen atoms of actinyl ion contain partial negative charge
and the effective charge on AnO22+ (An= U, Np and Pu) was found to be in between +3 to 3.3
complexation tendency rises with increasing atomic number due to increasing charge density
of actinide ions. Unlike lanthanides, actinides show preferences for soft donors also and
display covalent bonding behaviour with ligands containing soft donor atoms N, and S, which
govern the common basis for Ln/Ac separation. Actinide ions can form inner sphere (bonding
involve direct contact between cation and ligand) and outer sphere (ligand does not arrive the
primary coordination sphere and separated by minimum one solvent molecule) complexes
depending on the basicity of ligand. The outer sphere results into an exothermic and negative
entropy change while the inner sphere shows the endothermic and positive entropy change
during complexation of actinides. Actinides form complexes with π bonding ligands such as
alkyl phosphines and thiourea. In case of inorganic oxoanion ligands the stability of actinide
complexes varies in the order OH-, CO32- > F-, ΗΡΟ42- , SO42- > CI-, ΝO3- due to increasing
charge and basicity of the ligands. The actinides, because of their high positive charge, easily
hydrolysed via polarizing the water molecule sufficiently to release protons. For the actinides
in the +3, +4 and +6 oxidation states, the hydrolysis reaction occur in weakly acidic to
alkaline solution whereas in pentavalent actinides it takes place at pH ≥ 8. The early actinides
10
in +5 and +6 oxidation states exist as partially hydrolyzed actinyl ions, viz., MO2+ and
MO22+. Depending on the ionic potential and considering hydrolysis reaction akin to
acidic solutions with pH ≥ 0. The hydrolysis of actinyl cation decreases as UO22+> NpO22+>
PuO22+. Prior to complete hydrolysis, the actinides in hydroxo complexes tend to form
polynuclear species in which metal ions are allied through hydroxo (M-OH-M) and oxo (M-
O-M) bridges. The Pu4+ easily hydrolysed and form polymeric/colloid species (colloidal size=
1.5.4. Aqueous Speciation: The speciation of an actinyl ion is dominated either (i) by
hydrolysis, or (iii) by carbonate complexation depending about pH and the carbonate content
[16-17]
of the solution and also in some cases by both . Usually, the hydrolysis dominates the
speciation at the low pH values but above neutral pH, the carbonate (CO32-) present in the
solution works as a dominant ligand. The U-shaped solubility curves are observed due to the
development of negatively charged species, carbonate, and hydroxide, which increase the
solubility of actinyl as the pH increases above the neutral region. Even though hydroxo and
carbonato complexes dominate the actinyl speciation in many scenarios, however, there is
possibility that other inorganic or organic ligands can compete for the formation of actinyl
complex in pH range primarily deepens on its logβ value. Higher the value of logβ, larger
will be the pH range in which the actinyl-ligand complex can dominate over hydroxy or
carbonate. With the knowledge of a complete set of formation constants of all type of
and actinyl-hydroxy species, the speciation of actinyl ion can be described in a speciation
11
plot, which shows the proportion of each species as a function of pH, carbonate
concentration, or any other parameter. Fig. 1.4 illustrates the main complexes of uranyl that
of hydroxy and carbonate species of uranyl and neptunyl are described in Table S1.1. The
phenolates, phosphonates have a higher affinity for actinides and can consequently affect
their solubility and thus speciation. The complex formation (other than hydroxo complexes),
Figure 1.4. Aqueous species distribution of Uranyl in (a) absence (b) presence of carbonate
1.5.5. Molecular-orbital diagram: The actinide have unique feature of the linear ‘‘yl”‐ions
1.75 Å, which implies a strong multiple uranium-oxygen bonding. The orbital involved in
uranium-yl oxygen bonds formation is given in following molecular orbital (MO) diagram of
uranyl [Fig. 1.5], which shows that the bonding orbitals have mainly O 2p character and
turanyl(VI)-oxygen bond contains one σ and two π character thus it can be represented as
12
O2–≡U6+≡O2–. The bonding molecular
word borrowed from the biological sciences, has become a concept in analytical chemistry.
The biologist Orator F. Cook was the first to coin the term 'speciation' for the splitting of
[18]
lineages . The IUPAC has defined speciation as “species distribution based on their
(i) Chemical species: Chemical element: specific form of an element defined as to isotopic
13
(iv) Fractionation: Process of classification of an analyte or a group of analytes from a
certain sample according to physical (e.g., size, solubility) or chemical (e.g., bonding,
reactivity) properties.
its source, and its later fate is influenced by the physical and chemical properties of its
environment. The total element concentration may be static, the species may be highly
Metrical knowledge about metal ion’s speciation of the early actinides (U, Np, Pu and
Am) in aqueous (Aq) and non-aqueous (Non-Aq) solutions are of vital importance and
interesting to the nuclear industry in order to understand their solubility, stability, and
reactivity. One of the important features of early actinides is the formation of actinyl ions
(AnO2n+; n= 1, 2). Actinyl ions are the important species in the nuclear fuel cycle because of
[16, 19-21]
their multifaceted and highly characteristic chemical behavior . These ions have
dynamics behaviour in Aq solution and can form stable complexes and nanoclusters with
various organic and inorganic ligands in Aq and Non-Aq solutions. The chemical behaviour
i.e., speciation of actinyl ions, such as complex formation, dissolution, solubility, sorption,
precipitation, and immobilization, depend on several key factors of solutions chemistry such
first critical step for holistic understanding of early actinides, particularly actinyl ions, in
solutions. Metrical knowledge about the speciation of a metal ion under complexing medium
with varying physicochemical conditions (such as pH) can be broadly classified as below:
(i) Molecular speciation: It is the quest about stoichiometry, structure, chemical and stability
constant (logβ values) and thermodynamic parameters (enthalpy, entropy, and Gibbs free
energy) and coordination modes of a metal ion in complexing medium [Fig. 1.6].
14
(ii) Redox speciation: It deals with acquiring the information about the oxidation state
stability, redox energetic (peak potential and formal potential), trend towards reversibility,
parameters (αn and k0), electron transfer mechanism (ET; redox-mechanism) and
electrochemical phase diagram, i.e., Pourbiax diagram (Eh-pH plot) [Fig. 1.6].
actinides mostly depend on pH. For an example, in case of actinides, ET-mechanism may be
either step-wise or direct. It is also possible that ET coupled with some chemical reaction.
Speciation of actinides in solution is a daunting task and are highly challenging. This
is because (1) unlike most elements of the periodic table, there is absence of geologic
precedent, (2) their diverse chemistries, including rich redox behavior, radioactivity, and
[16, 19-21]
susceptibility towards hydrolysis and condensation reactions , and (3) the relative
stabilities of the actinide oxidation states and the types of complexes formed with the actinide
cations in these states vary for Aq and Non-Aq solutions. In addition, speciation of actinides
is very mobile and the effect of pH on their speciation is still couched. Based on aforesaid
cohesive database on their redox and molecular speciation. Therefore, novel research with the
15
aim of probing direct metrical information about molecular and redox speciation of actinyl
describing in-depth solution chemistry of early actinides and to mitigate challenges of NFC.
(i) Actinide environmental aspects for fate and transport of actinides (aqueous chemistry)
(ii) Dissolution, analysis, and recovery in non-aqueous medium (ILs and DESs)
(iii) Stability of unusual (or rare) oxidation state in aqueous and non-aqueous medium
(iv) Separation schemes advancement (oxidation states based or designing novel ligands)
radioactive contamination are nuclear weapons testing and production, processes associated
with the nuclear fuel cycle, accidental release of radioactive material used in medicine and
industry, mining and mineral processing, and the use of depleted uranium in weapons.
Besides the natural abundance in lithosphere (in Earth’s crust: U, 2−4 ppm; Th, 10−15 ppm)
and hydrosphere (2-4 ppm and 3 ppb, respectively), the light actinides (Th, Pa, U, Np, Pu,
16
Am, and Cm) are important environmental contaminants associated with anthropogenic
activities as discussed above. In the atmosphere, nearly 2х1020 Bq of radioactivity have been
[17]
released due to nuclear weapons tests . Accidental and other releases of radioactive
nuclides from nuclear reactors amount to 0.3% of this total. The primary releases consist of
residual fissionable nuclides in the weapon viz., uranium and plutonium and those which are
137 90
produced from the nuclear explosions (fission products, e.g., Cs, Sr etc. and nuclei
formed by nuclear reactions, beta decay, etc.). In the world’s aquatic systems, the nuclear test
explosions and reactor wastes have deposited an estimated 16х1015 Bq of plutonium [17].
1.6.1.2. Risks associated with actinides in environment: The spread of the actinides (or
[22]
their decay product) through aquatic pathways and physicochemical and geological-
mineralogical processes leads to significant environmental media (e.g., surface and ground
water resources and soil) contaminants, both in the immediate locality of the contaminated
sites and at a sufficiently long distance away from it. Their combined radioactivity (with a
longer half-life) and chemotoxicity are globally a serious concern for environmental chemists
since they pose chronic threats to human health (e.g., cell damage, kidney disease and lung
cancer) and environmental bio-diversity. In addition, the dissolved actinide ions are
[22]
potentially harmful to the aquatic ecosystem . Therefore, understanding the fate and
Aqueous speciation of actinides serves as the basis for (i) assessing actinides’ release
and transport rates i.e., mobility in the environment, (ii) long term safety assessment of deep
geological repositories, (iii) design methods for clean-up of contaminated site, and (iv)
chelation therapy for the development of formulations for the detoxication and excretion of
[16-17, 19-20]
actinides . This is because of the oxidation state and predominance of aqueous
complexes which govern their chemical behaviour and the interactions with the surrounding
17
geologic media in the environment. In addition, several important factors of solution
chemistry such as the solubility (hydration, hydrolysis, and solubility product), transport
composition, oxidation state, molecular-level structure, and nature of the phase in which the
about the chemistry which controls the interactions of actinides under environmental
[23]
conditions . The speciation of actinides critically depends on physicochemical conditions:
pH, Eh, metal and ligands concentration, nature of site-specific minerals and temperature, and
The overview of various geochemical processes that affects the chemical behaviour
and interactions of actinides in the environment are illustrates in Fig. 1.8. As evident from
[16-17]
figure, the important processes that may occur with actinides cations are (i)
precipitation, (ii) complexation reaction, (iii) redox process, (iv) sorption, and (v) colloid
formation, which are briefly discussed below. The dynamic interplay between these factors
18
affects partitioning of actinides between the solid and solution phases and retard or enhance
interaction between the actinides and environment fortifies for accurately assessing the
(i) Precipitation: It can limit solubility and have a retarding effect on release and migration
(ii) Sorption: It is the process through which actinides may sorbs on mineral or rock surfaces
in contact with the aqueous phase. Like the precipitation process, this process tends to retard
(iii) Colloidal formation: Actinides have a tendency to form or become associated with
colloidal size particle. In colloidal form, the actinides exhibit the dissimilar behaviour and
different migration than a dissolved species. This process can enhance or retard migration of
(iv) Redox process: The significant role of oxidation state and redox chemistry on the
migration of actinides in the environment is beyond dispute. The primary factor and the
single most important property of actinide ion determining the mobility of actinides ion is
[16-17, 20]
their oxidation state because (i) it can have wide range of values in aqueous phase
depending on the environmental redox conditions, and (ii) the tendency of geochemical
depends on oxidation state and differ from one oxidation state to another. However, for
actinides in the same oxidation state these processes tend to be similar. An interesting feature
of actinides is that their oxidation states can coexist in solution due to very small differences
in Gibbs energy between their oxidation states. It makes actinide redox distributions rather
difficult to predict. Since the chemical behavior of actinides depends mostly on the oxidation
state, actinide redox processes are of direct environmental relevance. Redox chemistry of
19
actinide under environment is reliant on the geochemical boundary conditions, where the
predominant actinide oxidation states and actinide species in aqueous solution can be
[16]
envisaged in diagram plotting the redox potential (Eh) versus the pH of solution . The
diagram is called as Pourbaix diagrams, which implies the expected dominant actinide
borders’ for aqueous systems and represent upper and lower limits for thermodynamically
stable aqueous solutions. The ability to predict actinide migration is directly linked to the
ability to establish the nature, stability, and mobility of dissolved species formed within the
bounds of the prevailing redox conditions. The environmental chemistry of actinides is vastly
dependent on their oxidation state distribution and associated redox conditions. The
surroundings near the nuclear waste repository are determined by the absence of oxygen (i.e.,
anoxic) in deep underground and presence of strong reducing condition. This type of redox
situation is mainly established by coupled direct and indirect microbial activity, geochemical
reactions with the host rock, and the presence of organic chelators, Fe(II) and S2- [24-25]. The
redox processes which influence the actinides oxidation states distribution in environment
(a) Abiotic: Reductants and oxidants: The anoxic environments contain a suite of potential
reductant Fe(II) species, S2- species and reduced natural organic matter (NOM) from which
oxidized species of actinide can accept electrons. The O2, Mn(III, IV) and Fe(III) species are
most abundant potential oxidants to whom reduced form of actinide can transfer electrons.
Radiolysis: The radiolysis of actinides in aqueous solution may also generate very reactive
species such as e-aq, H•, OH•, and H2O2, which can change the actinide oxidation state.
Redox effect of organic chelating agents with actinides: Organic chelators present in
environment from both natural and anthropogenic sources can directly affect the oxidation
state of dissolved actinide in two ways: (i) they can stabilize an oxidation towards
20
precipitation reduction by forming complexes and (ii) many organic chelators tend to reduce
state of the actinide and thus their speciation. Microorganisms mostly have an impact on the
redox behaviour of actinides in two important ways: (1) indirectly, by providing reducing
conditions in environment via the reduction of oxygen and by generating reducing agents (2)
bacteria DIRB and DSRB, respectively whereas microbial oxidation happen through
(c) Coupled abiotic-biotic redox processes: The microbial activity can produce a suite of
reductants/oxidants that can provide redox conditions for an example as reduction of Fe(III)
Actinides in oxic environment generally exists in higher oxidation states (e.g., UO22+, NpO22+
etc.) which are more soluble and mobile than their reduced or lower oxidation state species
(e.g., UO2 and NpO2 etc.). Lower oxidation state of actinide also exhibits enhanced tendency
more reduced/potentially less mobile form is an attractive approach for in situ immobilization
of actinide [24]. This kind of redox conversion can be obtained through above redox processes.
establish environmental fate of actinide. Redox behavior and thus above redox processes of
actinides can also be strongly influenced by electron transfer (ET) kinetics. In case of
actinides, the various possible ways of ET and their relevant nature of ET kinetics are
depicted in Fig. 1.9. As evident from figure, the redox reactions usually proceed fast which
involving only electron transfer (e.g., 1and 2) and no structural change. On other side, the
21
redox reaction requiring a molecular rearrangement either forming or breaking the actinyl
structure or coupling of chemical reaction with ET step are known to be slow and kinetically
Figure 1.9. Possibilities of electron transfer process and associated kinetic behaviour.
fulvic acids) or anthropogenic ligands (carboxylate or phosphonate based) (i) influence the
stability of redox forms and the solubility of actinides, and the actinide sorption processes (ii)
modifies their concentrations in aqueous solution and typically increases the actinide
solid phase, (iii) define the chemical form (speciation) of the actinides in solution, and (iv)
[16-17, 19-20]
control their presence in the environment . Thus, deep knowledge of the main
assessing sorption effects and modeling transport properties. Organic chelators, both naturally
[26-27]
occurring and anthropogenic act as “shuttles” for transporting actinides in the
22
[23, 26-27]
environment, play a pivotal role in determining their environmental behaviour .
Generally, the natural or anthropogenic organic chelators present near repository or mining
area have important role in the sorption/diffusion competition of actinides and can affect
sorption via redox and hydrolysis and diffusion via complexation with them. Both type of
organic chelators significantly influences the migration of actinide ion through strong
chemical interaction by forming aqueous soluble complexes and altering their redox energetic
and kinetics as a function of changes in denticity and coordination [20, 23, 28-29]. The migration
actinide ion through several environmental or synthetic processes, such as the formation of
adsorption/sorption complexes and precipitation by organic chelators [20, 23, 28-29], reduction of
higher oxidation state of actinide to their lower state, and microbial reduction process [20, 23, 28-
29]
. Actinide ion form strong bonds with organic chelators and its complexation process with
chelators tends to increase the stability of dissolved actinide species and slow down its
accessing the dependency of molecular and redox parameters on solution variables and for a
determining the fate of actinides in the environment. Research about exploring new organic
molecules that could provide an additional binding site to obtain selectivity and to prevent
Uranium and Plutonium are primarily utilized as a fuel source for nuclear power
production. Nuclear fuel used in power reactors consists of UO2 or mixed oxide (U, Pu)O2
(MOX). Once discharged from the reactor, the spent nuclear fuel contains unreacted U/Pu
[32-33]
oxide, fission products, and transuranic (TRU) elements . For a standard PWR with a
burn up of 33 GW/t, the spent nuclear fuel typically consists of about 95.5% UO2 matrix,
23
[32-33]
3.6% fission products, and 0.9% transuranium elements . The increasing demand of U
and Pu as fuel materials in the nuclear reactor and also for the sustainable development of
nuclear energy, the processing and reprocessing of these metals from naturally occurring ores
and spent nuclear fuel (SNF) is of great importance. Current approaches of reprocessing fall
into two major categories, aqueous based hydrometallurgical and pyrochemical processes [34].
Reduction Extraction (PUREX) process, the spent nuclear fuel will be dissolved first with
high concentration HNO3 and then solvent extraction carry out for the separation of U and
Pu. This step results in a large volume of high-level radioactive liquid waste with most of the
fission products and minor actinides. The long-lived fission products and minor actinides
pose potential long-term risk to the environment. Further, the principal disadvantage of
PUREX is in the area of proliferation resistance [33-35], as this process permits the isolation of
a pure plutonium product. In pyrochemical reprocessing method the actinides are separated
[35]
from the high temperature molten salt media through electrochemical means . However,
the requirement of high temperature and the strongly corrosive nature of the molten salt
media have been opposed critically for its application in reprocessing. Further, both the
processes have poor atom and energy efficiency. Thus, complexities of both processes and
International Atomic Energy Agency (IAEA) safeguard policy is based on the robust
nuclear materials accounting (NMAC) which ensure no misuse of nuclear materials (NMs) or
[36]
technology for nuclear weapons . The environmental swipe samples at a declared facility
are collected and transported to analytical laboratories for dissolution and estimation of NMs
for proper implementation of safeguard. Furthermore, for quality assurance of reactor fuels
and nuclear forensics studies, the U and Pu samples (including swipe samples) were routinely
24
preparation is time consuming and hazardous dissolution procedures, which uses corrosive
inorganic acids (e.g. HNO3, HF, HClO4, and/or H2SO4) [37-38]. For operator and environmental
safety considerations, such corrosive chemicals should be avoided because it generates large
volume of harmful NO and NO2 gases and increases neutron and γ-ray background due to (α,
nγ) reactions on 19F nuclei. Based on above issues, for advancement of NFC, it is necessary
Dissolution of metals or their oxide/salt is an important and challenging issue for the
development of alternative route in the nuclear field. The direct dissolution of metal oxides
has the added advantage of complete mitigation of the aqueous processing of SNF. If a green
present and advanced fuels as well as in routine sample dissolution at laboratory scale. It will
liquids (ILs) or novel solvents such as deep eutectic solvents (DESs) are developed, which
[33, 39-43]
can work at room temperature itself . Non-aqueous routes are talented new
technologies for NFC since it has several advantages such as minimal volumes of waste
generation, less time consuming, low criticality concern and feasibility to reprocess the high
burn-up, short cooled fuels compact infrastructure, radiolytic stability of the process reagents
and ease of recovery of the metals by electrochemical means etc. If ionic solvents such as ILs
or DESs are involved to process metals or their oxides/salts then the term “ionometallurgy”
can be used. In ionometallurgy, an innovative and more efficient ways of processing and/or
reprocessing and the subsequent production of metals can be achieved via simple two-step
or electrowining process) without generating hazardous wastes [39] [Fig. 1.10]. Dissolution of
25
metal oxides depends on many factors, such as temperature, mixing time, lattice energy,
crystal structure, acidity or basicity of the metal oxides, stability of the metal ion complex
[43]
formed after dissolution or by the presence of chloride ion in ionometallurgy . For a
complete dissolution, the lattice of the solute particles is required to be broken to become
solvated by the solvent molecules. For this, the attractive force among dissimilar particles
(solute-solvent) should be higher than the attractive force between similar particles (solute–
[33]
solute and solvent–solvent). Fan et al. proposed a correlation (U/x) between the lattice
energy (U) of a metal oxide (MxOy) and its solubility in the IL medium. A good solubility of
metal oxides was found if U/x < 7000 kJ mol-1 whereas if U/x >10 000 kJ mol-1 metal oxides
is insoluble. In viscous ionic liquids the dissolution process of metal oxides is slow due to
slow mass transfer. Room temperature ionic liquids (RTILs) were put forth for the purpose of
[39-43]
nuclear fuel reprocessing and other related campaigns . However, their potential was
never realized due to issues such as cost, toxicity, and non-biodegradability. In addition, ILs
[39, 44]
(and molecular solvents) have poor solubility for metal oxides because of (1) high
strength of metal oxide bond, (2) the inability of the oxygen anion of a metal oxide to exist in
a free form in ILs, (3) the absence of the oxide-binding reagents in ILs medium to react with
the metal oxide, and (4) weak coordinating ability of constituting anions and cations.
Figure 1.10. Scheme of the general idea of metal oxide processing in non-aqueous medium.
26
Task-specific ionic liquids (TSILs), which are ionic liquids with a functional group
covalently tethered to the cationic or anionic part, are developed to overcome the problems of
[44]
poor solubility of metal oxides . However, the main drawback of TSILs is that they are
often only accessible after a difficult and time-consuming multistep synthetic procedure
[44]
which restricts their use in large-scale industrial applications . Meanwhile, deep eutectic
solvents are the newest entries to the current world of chemical media with unique
physicochemical and tuneable properties. It is innovative that DES can be explored for the
dissolution and recovery of actinides from their primary, secondary sources, spent fuels or
1.6.2.2. Speciation of Actinides in non-aqueous medium (ILs or DESs): One of the key
chemical aspects when considering the reactivity of a metal during ionometallurgy is the
species that form upon dissolution [45-49]. The speciation of metals in ILs or DESs solution is
extremely pertinent in the area of metal salts/oxides dissolution and it basically controls their
reactivity. A species dissolved in these solvents will experience a very strong ionic
atmosphere since both consist only of ions. It will be quite interesting to understand how the
chemistry of metal ions behaves under such an extraordinary condition. Speciation is known
to regulate key features of metal processing and electrochemical behaviour, such as solubility
and redox potentials. In aqueous solvents, the chemistry of H+ and OH- dominate the species
formed in solution, the redox properties, and the solubility of metals whereas in ILs and
DESs the speciation is mainly governed by the anionic component of the liquid. Metal ion
interacts with whatever ligands (L) are available in ILs or DESs (equation 1.1). The species
present and the equilibrium constant are depending upon the concentration of salt in solution
27
The solvent species which dominates complexation is the anion because it has a significant
Lewis basicity and it is generally present in higher concentration (5-10 mol dm−3). It means
For foreseeing the application of ILs or DESs as media to the reprocessing and waste
about the both kind of speciation of actinides in ILs or DESs is essential to better understand
the chemistry that is occurring starting from their dissolution to subsequent selective recovery
dissolved actinide metal ions is imperative (1) to benchmark the redox thermodynamics and
kinetics of redox-sensitive actinide ions and their dependency on solvation and equatorial
coordination, (2) to describe the coordination environment around the actinide ion, (3) to
reveal the coordination competition between the anionic/hydrogen bond donor ions of ILs or
DESs toward equatorial coordination of actinide ions, (4) to provide the optimized redox
conditions for actinide ions analysis and electrocatalytic recovery, (5) to understand the
activity of actinide ion in ILs/DESs solution which is crucial for utilizing the full potential of
a reactive species, and (6) to understand the actinide oxides/salts dissolution phenomenon
such as rate, kinetics and mechanism. The objective of the redox speciation of actinides in
ILs/DESs is to establish the baseline thermodynamic and kinetic data for these systems which
will be helpful to critically evaluate the ability to use electrochemical methods and ILs/DESs
electrodes surfaces [49]. Redox speciation studies will examine the interfacial electron transfer
processes of actinides and their potential dependent deposition at electrode surfaces. The
28
medium is expected due to difference in speciation of the metals in the two solvent systems.
The solvation plays an important role in determining the complexation and coordination
behaviour of metal ions in ionic liquids or DESs medium and provides direct information
[45-48]
about the speciation , which will be helpful to reveal the mechanism involved in
behaviour, and to know about the feasibility of cation-cation interactions between homo or
hetero actinyl ions. The solvation of actinide ions in these solvents is different from what is
observed in conventional organic solvents and water since the poorly solvating behaviour of
ILs/DESs can provide the formation of coordination compounds with low coordination
numbers. It is of great importance to study the solvation of metal ions and their molecular
thermodynamic parameters such as stability constant, enthalpy, and entropy of actinides. All
SNF reprocessing media is quite a challenging analyte due to its high radioactivity
and strong acidity, from the analytical chemistry point of view. Speciation characterization of
the actinide elements extracted and their relative concentration determination is crucial for
the monitoring the performance at each step of the reprocessing and chemical quality control
of nuclear fuels, ensuring security of the nuclear fission processes, and measuring the nuclear
wastes. In addition, in order to implement full scope of lAEA Safeguards under an agreement
between the nuclear non-proliferation treaty signatory state and the IAEA (INFCIRC/153;
IAEA Safeguards Glossary [36, 50]), robust nuclear materials accounting and control (NMAC)
and nuclear forensic are essential for (i) precise and accurate determination of nuclear
29
materials (NMs) and (ii) analysing swipe samples at both bulk and trace/ultratrace level
concentration to detect and deter any unauthorized activities and to verify the identity of
smuggled or discovered nuclear materials [50]. It is generally accepted that trafficking of metal
ions/metal complexes depend greatly on their actual chemical and physical forms, oxidation
state, etc., and not merely on their concentrations. Thus, it is necessary to determine not only
the total content of the element but also its individual chemical and physical form or the
oxidation state, i.e., speciation information, which can be attained through extensive
methods
Destructive assay (DA) provides a more accurate quantifiable measurement than non-
destructive assay (NDA) techniques and can be used to determine if a small fraction of a
[51]
material is present . In most cases, the precision of DA techniques is better because the
effect of the sample matrix can be eliminated or corrected for. Detection limit of DA
techniques is generally lower than NDA techniques since the measurement approaches
mainly inductively coupled plasma atomic emission spectrometry (ICP-AES) and mass
spectrometry (ICP-MS). These methods are rather time-consuming and can barely be
implemented in on-line mode. The results of such analyses are usually available only several
hours after sampling. Other optical spectroscopy tools suffer from the incapability to deal
with non-transparent and highly scattering media and also cannot ensure sufficiently low
detection limits for NMs. Conventional electrometric redox titrimetry using endpoint
30
used for analysis of actinides/NMs. However, these conventional methods have several
disadvantages, viz., generation of hazardous acid waste, the involvement of various metallic
impurities (Cr, Fe, Ti, Ag, etc.) due to the addition of redox titrants during the analysis, time-
consuming hazardous acid dissolution, and the laborious recovery of actinides/NMs from
[52]
analytical waste . Coulometry is an absolute electroanalytical method, but the total period
of analysis for a sample is effectively long. Several studies were further devoted to
modification of electrode surfaces for NMs analysis as an alternative but its preparation
involves a complex procedure and is expensive, tedious, and time-consuming, which restricts
[53]
its application in routine sample analysis at bulk concentration . Side effects of using
conventional analytical methodologies in routine chemical analysis may involve serious risks
for operators and damage to the environment due to used toxic reagents. Therefore, new
analytical methodology which consider the use of less toxic reagents and minimizes wastes
generated need to be developed in order to determine NMs with good precision and accuracy.
Among the analytical techniques, electroanalytical presents many advantages for the
simple, cost effective and rapid electroanalytical techniques capable for fast, precise, and
accurate destructive determination of concentrations of the analyte [54]. It has twin advantages
as the possibility to work at larger concentration ranges (from milli to micro molar) and the
inherent sensitivity towards oxidation states of the elements. It also enhances the possibility
of on-site analysis, ability to design the facilities to significantly lower the radiation exposure
to the work personnel and equipment and is inexpensive, enabling its wider application. The
technique involves minimal sample handling, thereby reducing the potential sources of
contamination. Because of its wide concentration range, it is applicable for (1) analysis
31
concentration for actinides/NMs. The electrochemical sensing of actinides/NMs appears to be
target analytes and allows measuring in turbid liquids. The measurement of the redox
through the formation of Pt-O and Au-O bond [55], (2) side reactions such as water oxidation,
metal oxide formation, hydrogen evolution at negative potentials etc., and (3) much narrow
cathodic potential window of Pt and Au. In most cases, the reduction potential of actinides is
significantly more negative than the potential associated with the hydrogen evolution side
[56]
reaction is ~ -0.2 V . Because of these factors, the analysis, especially the simultaneous
has wider potential window, thus analytical activity such as simultaneous determination of
[57]
NMs is quite feasible in non-aqueous medium. Literature reports indicate that DESs can
be use as designer solvents for green analytical chemistry because the intermolecular
interactions exist between supramolecular structure of DES and target analytes. DESs
behaves like liquid crystals in which the solute molecules are fixed in crystals, therefore,
analytical detection enhancement is possible when target analytes are part of DES (a
Uranium (U) is known to exist in various oxidation states as III, IV, V and VI. Among
them pentavalent uranium (i.e., UVO2+) has historically been reported as unusual (rare)
(V)." Chemical Reviews 1969, 69, 5, 657–671 [58]. It remains comparatively unstable and rare
due to (1) its propensity to disproportionate to thermodynamically favoured U(IV) and U(VI)
32
products in aqueous solution according to following equation 1.2 and (2) its extreme air and
water sensitivity and its easy conversion/oxidation to hexavalent uranium in the presence of
Unlike NpO2+ and PuO2+, which are stable under a wide range of conditions,
pentavalent uranyl can be stabilized in concentrated carbonate media or at low pH (pH 2-3)
[59-60]
, where in each environment it exhibits a considerably longer half-life [2], but otherwise
obscure. Recent kinetic and theoretical studies suggest that the redox disproportionation
reaction proceeds through the interaction of the “yl” oxo group of UO2+ with an adjacent
pentavalent uranyl, forming cation-cation complexes (CCIs) that facilitate electron transfer
and yield UO22+ and U(IV) species after protonation of the uranyl oxygen atoms. The CCI
complex mainly contains two mutually coordinated UO2+ groups in which -O of one UO2+
moiety coordinated to U of second UO2+, as shown in Fig. 1.11. Its formation (or interaction
33
between two pentavalent uranyl moiety) is expected to favoured by the increased Lewis
basicity of oxo group (U=O) in pentavalent uranium (UO2+) than to that of hexavalent uranyl
(UO22+). Previous studies [59-60] reported that CCI complex mainly exist in two forms either in
units are assembled through a diamond shaped via two mutually coordinated monodentate
UO2+ group. The value of the U=O distances are 1.941 and 1.850 Å while the O(1U)-U(1)-
O(2U) and O(1U)-U(1)-O(2U) angles are 175.80º and 101.6º, respectively. The structure of
coordinated to each other in a monodentate fashion to form a square plane. Each UO2+
coordinates two adjacent uranyl groups and is involved in two T-shaped cation-cation
interactions (two linear UO2+ groups arranged perpendicular to each other). The distances of
U=O bonds in structure of above complex are mentioned in Fig. 1.11 while the angle are as
the study on solution chemistry of pentavalent uranium is difficult. Thus, gaining information
about aqueous speciation of pentavalent uranium is highly challenging, resulting in its elusive
aqueous chemistry and a knowledge gap related to the comprehensive understanding of its
stability and structure in the aqueous solution. The aqueous speciation of pentavalent uranium
plays a crucial role in the solubility and mobility of uranium in the environment and in
investigating the mechanism of biotic and abiotic reduction of uranyl species in the
environment. Despite of its huge instability, the pentavalent uranium can be stabilized against
disproportionation under specific conditions by careful choice of ligands and chemical media,
which prevents the formation of CCI, such as equatorial complexation or uranyl oxo binding
by (1) sterically bulky organic ligands/chelators viz., shiff base ligands (salophen) or
counterions such as cation K+ or iron, which provides the steric protection along the O=U=O
34
[59-60]
axis , and (2) anaerobic environment and non-aqueous solvents. Due to rapid
the aqueous speciation of UVO2+ species in comparison to other techniques since it allows for
Separations of actinides are a critical aspect to reprocess the spent fuel for subsequent
safe disposal and for nuclear waste management. This can be achieved either by (a) control
and determination of actinide oxidation states or (b) designing novel ligands for complexation
of actinide species.
(a) Oxidation states play a considerable role in actinide separation especially in case of early
actinides (U, Np, Pu and Am) since these exist in more than one oxidation states and every
of actinides in solution and thus it is feasible to easily partition/separate one state from
another. (b) In actinide separation chemistry, designing novel ligands for selective
complexation of actinides is an important task. This research area has gained significant
attention due to their potential applications in nuclear waste management. The newly
designing ligands should be talented to enhance the binding affinity, specificity and
preorganized macrocyclic ligands (e.g., Cucurbituril CB[n], and cyclodextrin) and chelating
several ligating unit on a molecular platform provide a specific fit to the target metal, which
results in a better extractant with higher metal selectivity. The chelating ligand with
35
stereognostic coordination platform has a tendency to form an intramolecular hydrogen bond
(IMHB) with the oxo group of the actinyl ions and imparts specificity towards actinyl ions
binding [30-31], which helps in molecular recognition and selective separation of actinyl ions in
presence of other interfering ions. To develop and optimize new separations chemistries,
The principles about choice of chelating ligands, chemical media, and ligand
designing should be depending on their relevant application and important role in mind so
that new research activities can assist in actinides’ separation, geosphere migration
predication, dissolution and analysis, redox control, and actinyl-oxo bond (i.e., O=U=O)
activation. Keeping these in mind, in the present dissertation, the chelating ligands and
chemicals media, as discussed below [Fig. 1.12], have been prudently chosen to cover all
speciation aspects discussed under title (i) to (iv) and their respective “redox and molecular
speciation” data with “actinyl ions” were generated. The details about the aqueous based
ligands and DES media, used in present thesis, are given below. Phosphorus oxo anion and
carboxylate-based ligands were selected due to their high aqueous solubility and high
(i) Citrate: Citric acid is a naturally occurring a-hydroxy polycarboxylic acid. It is capable to
form strong water-soluble complexes with actinides over a wide range of pH.
which shows stronger complexation tendency towards the metal ions than the isostructural
therefore metal ion can more effectively compete with H+ for truly anionic ligand binding
36
Figure 1.12. Choice of complexing agents
moiety has been changed by a carboxylate group. It has high complexing strength for metal
(iv) Deep Eutectic Solvents (DESs): DES, a new class of ionic liquid first introduced by
[62]
Abbott et al. in 2001 , constitute a
the 21st century. DES is generally composed of two or three cheap and safe components,
generally one component as hydrogen bond acceptor (HBA) and other as hydrogen bond
37
donor (HBD), which are capable of associating with each other through hydrogen bond
interactions to form a eutectic mixture with a significant depression in freezing point than that
[63-64]
of each individual [Fig. 1.13] . The term DES used to differentiate it from ionic liquids
which contain primarily one type of discrete anion and cation. In contrast to ILs, the DES
contains a variety of anionic and/or cationic or neutral species. DESs that are made of
primary metabolites such as organic acids, amino acids, sugars, polyols, and choline
derivatives are entitled as Natural Deep Eutectic Solvents (NADES). NADES fully represent
green chemistry principle. DESs have the general formula as Cat+X-zY, where Cat+ is in
whereas X is a Lewis base (usually a halide ion). Y is the Lewis or Bronsted acid and z is
number of Y molecule which can interact with anion X- and form complex anionic species
[45]
. The classification of DES is represented in Table 1.2 and some examples of HBA and
The physicochemical properties, advantages, and applications of DESs are described below:
38
Figure 1.14. Structures for several HBAs and HBDs for DES synthesis.
(i) Phase diagram: The word “eutectic” comes from the Ancient Greek εὔτηκτος
τηκτος or eútēktos
which
ich implies easy melting. DESs are not pure compounds but mixtures of two or more pure
(∆Tf) is the difference between the ideal eutectic temperature of the composition of a binary
mixture of A + B and the linear combination of the melting points of the pure components A
and B, as shown in Fig. 1.15,, and further stated that the magnitude
magnitude of ∆T
∆ f is directly
proportional to the strength of interaction between A and B (larger the interaction the larger
[65-66]
will be the ∆Tf). However, some recent studies indicate defining a DES is still
mixture of two or more pure compounds for which the eutectic point temperature is below
39
that of an ideal liquid mixture, presenting significant negative deviations from ideality
(∆T2>0)”, where ∆T2 stands for depression in the temperature which is the difference
between the ideal and the real eutectic point [Fig. 1.15]. Furthermore, they suggested that a
DES should be liquid at operating temperature even if this needs a different composition than
the eutectic one. The above definition by Martin et al. provides the appropriate meaning of
word “deep” in DES as “huge depression in freezing point” and covers the mixtures with a
melting point lower than the ideal eutectic temperature. The magnitude of the freezing point
depression, ∆T2, is defined as Tf(real) – Tf(ideal), where Tf(real) is the measured freezing point of a
mixture at the eutectic composition and Tf(ideal) is the theoretically predicted freezing point for
an ideal mixture. In DESs formation, the significant melting point depression of a eutectic
mixture compared to that of the pure materials is because of charge delocalization from the
Figure 1.15. Phase diagram of a two-component mixture representing the eutectic point
(ii) Freezing Point (Tf): DES generally has a lower freezing point than that of individual
interaction between the anion of HBA and the HBD component. Abbott et al. suggested that
that the freezing point of HBD-salt eutectic mixtures mainly dependent on (1) the lattice
40
energies of DESs, (2) the mode how the couple anion-HBD interacts, and (3) the change in
entropy due to the formation of a liquid phase. The selection of HBA and HBD is the key
factor which decides the lower freezing points. For a chlorine salt-derived DES, the freezing
point decreases in the order F- > NO3-> Cl- > BF4-. The molar ratio of HBA and HBD has also
a noteworthy impact on the freezing point of DESs. In carboxylic acid based DESs, in order
to form eutectic mixture, two carboxylic acid molecules are required to complex each
chloride anion of choline. In the case of diacids, the eutectic is formed at its 50 mol% (i.e.,
1:1). In addition, the carboxylic acid with the lowest molecular weight shows the largest
depression in the freezing point. The freezing points for all reported DESs are below 150°C
and the DESs with a freezing point < 50°C are more applicable for various field.
3. Density: The majority of DESs show higher densities than water with the value ranging
from 1 to 1.3 g/cm3 at 25°C for most of the DESs while metal based DESs have densities in
the 1.3-1.6 g/cm3 range. The difference in the density of DESs might be due to a different
vacancies. During the DES formation, the average holes radius was decreased which results
in a slight increase of the DES density compared to that of neat HBD. The mole ratio of
HBA/HBD, and the nature of salt (HBA) and the HBD have an important effect on the
linearity with the increasing temperature. Lower densities than water are obtained for
hydrophobic DESs.
4. Viscosity: Most of the reported DESs exhibit relatively high viscosity (>100 cP) at room
temperature except choline chloride (ChCl)-ethylene glycol which has low viscosity as 37 cP.
Sugars or carboxylic acids based DESs shows extremely large viscosity (12730 cP for 1:1
ChCl: sorbitol) and even high viscosities were reported for metal salts based DESs (85000 cP
[67]
for 1:2 ChCl:Zinc Chloride) . The low viscosity is observed only in hydrophobic kind of
41
DESs. The viscosity of DESs is mainly governed by (i) forces such as hydrogen bonds,
vander waals and electrostatic interactions, (ii) the chemical nature of DES components
(HBA and HBD), (iii) the mole ratio of HBA and HBD and (iv) the temperature and water
content. The high viscosity of DESs is mainly due to extensive hydrogen bond network
between each component, which consequences in lower mobility of free species within the
DES. In addition, the large ion size and very small void volume of the DESs give the high
5. Ionic conductivity: The majority of the reported DESs have low ionic conductivity
compared to high temperature molten salt (7.61 mS/cm in ethaline compared to 350 mS/cm
in BiCl3 at 200°C). However, their conductivity is comparable and/or higher than drinking
water (0.2 to 0.8 mS/cm) and deionized water (0.05 µS/cm), respectively. With raise in
of the DES viscosity. The conductivity of DESs is also depends on the HBA/HBD mole ratio,
the nature of the organic salt and the hydrogen bond donor as well as the salt’s anion, and the
addition of water.
6. Acidity and alkalinity: The chemical nature of HBDs has a strong effect on acid and basic
strength of the DESs. The Hammett function (H_) is used to evaluate the acidity and basicity
of non-aqueous solvents. The H_ value of Reline is 10.86, which suggest that this DES is
weakly basic. The Maline (1:1) is acidic DES with pH 1.67 and its acidity varies with
addition of water such as for Maline:0.22 H2O, the pH lies in the range 1.28-0.41.
7. Effect of Water: When water is added to the DES, at a low water content, it integrating
into the deep eutectic solvent’s network and strengthening the hydrogen bonds taking place
[66]
between the HBA and the HBD . Further dilution of DES will result into hydration of all
42
the species; however, it is anticipated that the anion will be preferentially hydrated. Water
8. Advantages of DES and Major Applications: DESs are environmental benign and have
low ecological footprint due to their several advantages such as non-volatile, non-toxicity,
inertness with water. DESs are less expensive and chemically tuneable and their synthesis
such as electrochemistry,
metal/metal oxides
processing, catalysis,
extraction/separation, carbon
[45]
dioxide adsorption, cleanup of heavy metal and nanotechnology . The foremost
applications for DESs include the incorporation of metal ions in solution. DESs are talented
of donating or accepting electrons or protons which confers them excellent metal ion
hijacker [Fig. 1.16], especially Maline, due to high solvating potential for metal oxides/salts.
DESs have wider potential windows than aqueous solutions and high electrical conductivity,
According to IUPAC Green chemistry is “the design chemical products and processes that
43
[68]
plants, and the environment” . Green chemistry basically is based on twelve principles
The OECD in 1990 define the term Sustainable chemistry as “a scientific concept that seeks
to improve the efficiency with which natural resources are used to meet human needs for
[68] [69]
chemical products and services” . In 1999, Hutzinger et. al provided a fundamental
difference between these two concepts of chemistry and suggested that the green chemistry
denotes to the “design, manufacture, and use of chemicals and chemical processes that have
concept of sustainable
Sustainable chemistry”
terms. The essentials of Green and sustainable chemistry are (i) Holistic approach and long-
term positive development, (ii) validation of 12 principles of green chemistry, (iii) address
social, environmental, scientific, and economical aspects, and (iv) substances and process
44
Speciation of uranyl in citrate is very mobile and shows complex behavior about pH-
Despite the extensive spectroscopic studies, no organized study has been carried out on the
speciation of UO22+ in citrate medium considering its redox behavior. In addition, despite
decades long research as many discrepancies among published data exist and knowledge gaps
remain. In aqueous solution, the speciation of uranyl with PPA is unknown in literature due
sensitive technique which can work at trace level concentration of uranyl. In this context,
complexation studies on Np have been reported with inorganic ligands (CO32-, ClO4-, NO32-
and SO42-) and some other organic chelating ligands including carboxylic acids, amides, and
hydroxamic acids. However, no reports in the literature have been found on aqueous solution
no studies have been reported on aqueous complexation, redox speciation, and coordination
viz., PFA, PAA and 3-PPA. Literature survey implies the unavailability of redox
thermodynamics and kinetics parameters of uranyl and Np with citrate, PPA and
speciation of UO22+ with varying physicochemical conditions in aqueous citrate, PPA and
[39, 70]
phosphonocarboxylates and Np in PPA. Previous studies reported the use of additive
viz. water, aqueous hazardous acids and Fe/Al-salt during ionometallurgical processing of
uranium (U) matrices. Literature survey implies the lack of direct sequestration of uranium
45
(U) from its solid matrices and the greening analysis methodology for U. Electrochemical
studies on d-block metal ions [Fe(II)/Fe(III)] and Cu2+/Cu+ in Reline and Ethaline were
actinides in DESs. In addition, the important aspect that how DES interact with uranium is
missing in literature. Use of DES as co-solvent for Am separation [71] is available in literature
fundamental chemical aspect (1) to understand their behaviour such as redox potential and
solubility, (2) to predictive their migration in the aquatic environment and associated risk
assessment for nuclear waste disposal, and (3) for advancement of new analysis, separation
and recovery schemes based on non-aqueous methods. The migration of actinides within the
chelators, redox reactions, and physiochemical conditions (Eh and pH). Aqueous soluble
organic chelators, both naturally occurring and anthropogenic in the environment, play a
pivotal role in determining the migration behaviour of actinides. Carbon and phosphorous are
two most important elements in the bio-geosphere as they are omnipresent in both biotic and
abiotic medium. Phosphonate and carboxylate are constituents of the cellular wall of bacteria
and considered as building blocks of glyphosate and humic substances. These are the main
driving functionalities for most of the chemical interactions associated with the actinides in
the aquatic environment. Deep eutectic solvents with tuneable chelating functional groups are
a suitable sort of “green and sustainable” non-aqueous media and works as a “designer
solvent” in analytical chemistry. It can provide high solvating potential for actinides solid
matrices to leach their ions from oxides, salts, or alloy host core. DES, especially type III,
46
provides a new variety of organic complexing moiety, which is different from organic
complexing ligands in aqueous solution. The malonic acid-based DES such as Maline, which
contains dicarboxylic acid functional group, is strongly acidic and its solvating potential for
metal oxides is correlated well with that in aqueous HCl. The imidazolium cation is quite
stable to radiolytic degradation and its density, surface tension, refractive index, etc. are not
notably affected upon irradiation. The present dissertation aims to acquire comprehensive
redox and molecular speciation of early actinides with aqueous and deep eutectic solvents
based complexing agents. To accomplish the aims, the present thesis has mainly two
(a) Aqueous based studies using complexing ligands: Speciation of actinides in presence of
aqueous soluble carboxylate and phosphonate based chelating ligands. In this context, citrate,
actinides): In the present studies, Maline and imidazolium cation based DESs were preferred
understand the mechanistic of dissolution process, interaction of uranium with DES and
analysis parameters.
Attempts have been made to obtain the redox potentials (peak and formal potentials),
Pourbiax diagram (Eh-pH plot), transport property (i.e., diffusion coefficient), electron
transfer (ET) mechanism, heterogeneous electron-transfer kinetic parameters (k0 and αn),
47
wave voltammetry, chronoamperometry and chronopotentiometry were employed to probe in
details (1) the redox speciation of actinides with varying pH in aqueous and DESs, (2) to
inspect the coupled chemical reaction with electron transfer steps, and (3) to obtain the
calibration plot for analysis of uranium and the stability constant value by Deford-Hume
coupled chemical reaction, ET-mechanism, and stability of pentavalent uranium (UO2+) and
to acquire its electronic absorption spectra. UV-visible absorption studies were carried out to
obtain the stoichiometries and their respective stability constant values. The fitting of
absorption titration data was carried out with Hyperquad software. Potentiometry studies
aqueous. The coordination modes of the complexes were analyzed by EXAFS. MINTEQA2
and HySS software were used to obtain the speciation diagram as a function of pH. Physical
properties, such as density, viscosity, and melting point of imidazolium-based DESs were
determined. Karl-Fischer titration was used to determine the water content in DESs.
Theoretical calculations viz., molecular dynamics (MD) and density functional theory (DFT)
were performed to (1) authenticate the experimental results, (2) optimize the geometries of
the dominant species, and (3) derive the bond distances, partial charges, and molecular orbital
(MO) diagram for the optimized geometries. The studies in present thesis do not cover the
isolation of pentavalent uranium. Herein, the speciation studies of actinides are limited to
aqueous and DES solutions. The solid-state characterization of actinide complexes and
speciation of actinides in other non-aqueous media such as ionic liquid are not within the
48
1.11. Chapters’ distribution Present dissertation has been divided in six chapters as
49
Chapter-2
AIM: Gaining comprehensive information about redox and molecular speciation of actinyl
ions in aqueous and non-aqueous solutions cannot be feasible simply through a single
about the various techniques used for speciation studies and sample solutions preparation of
actinyl ions, organic ligands, dissolution process of actinide matrices and synthesis of deep
50
2.1 Introduction: The studies in the present dissertation are mainly focused on
understanding the speciation of actinyl ions with aqueous soluble chelators and deep eutectic
solvents. The acquiring speciation information of actinyl ions in aqueous and deep eutectic
(i) Preparation of metal ions and organic chelator solutions in aqueous solution and
their pH optimization: Under metal ions, the preparation of uranium and neptunium
solutions is described. The organic chelators employed are citric acid, phenylphosphonic acid
and three phosphonocarboxylic acids. The preparation of above organic chelators in aqueous
(ii) Synthesis of deep eutectic solvents and dissolution procedure of actinide matrices:
The synthesis of Maline, two imidazolium based DESs, Choline dihydrogen citrate-Malonic
acid DES are described. Further, the dissolution procedure of uranium matrices (oxides, salts,
depth understanding of
discussed.
51
(i) Aqueous media: AR grade sodium perchlorate, PPA, Citrate and three
phosphonocarboxylates (PFA, PAA and 3-PPA) were used to carry out the experiment. The
stock solution of uranyl nitrate (0.1 M), prepared by dissolving the solid U3O8 in 1-2 M
HNO3, and was standardized using Davies and Gray’s method. Samples were prepared by
mixing stock solutions of UO2 (NO3)2, NaClO4, and ligands to obtain solutions with required
NH3 using PICO+ pH meter which is based on a combination pH electrode (glass membrane
electrode and a reference electrode) with a glass electrode (Lab india, Mumbai, India). The
pH meter was calibrated using standard buffer solutions (Merck) of pH 7, 4 and 9. All
solutions were prepared using milliQ water having resistivity 18 MΩ·cm. ESI-MS
measurement was done in 100 % Methanol. Supporting electrolyte was used as 0.1 M
NaClO4 in the electrochemical experiment. Np(V) Stock Solution: Np-237 stock solution in
our laboratory was used for the present studies. It was purified from Pa-233 and other cations
[72]
using anion exchange (Dowex 1X4) 57 resin in nitrate form . The acidity of the Np stock
solution was adjusted to 7.5 M (HNO3) for loading on the anion exchange resin (Dowex
1X4). To this solution, ferrous sulphamate and hydrazine solutions were added to ensure that
all Np(VI) and Np(V) is reduced to Np(IV). This feed solution was loaded on the column and
after three column volume washings of the resin with 7.5 M HNO3, the solution was eluted
with 0.35 M HNO3 to collect the purified Np solution. The purity of neptunium was checked
(ii) DES media: (a) Maline and Choline-Dihydrogen Citrate-Malonic acid DES
synthesis: Choline chloride (≥98%) (CC) and Malonic acid (99%) (MA) were purchased
from Sigma-Aldrich. Both solids were taken in 1:1 molar ratio and mixed well and kept for
heating at 60oC for four hours till the homogenous Maline deep eutectic solvent formation.
52
METALAB oven was used for heating the solid samples. The karl-fischer titration was used
to determine the water content in prepared MALINE, providing nearly 1% of water content.
clear liquid at room temperature (RT) but appears as gel-type was prepared following the
[74]
procedure given in the previous literature report . Malonic acid (MA) and diglycolic acid
clear single viscous phase was obtained [Fig. 2.2]. Hereafter, the DES C10mimBr with MA is
referred to as DES 1 and that with DGA is referred to as DES 2. The amount of water content
with drying both DES 1 and 2 were found to be ~ 8% and 6.7%, respectively.
(c) Source of Uranium matrices and their Dissolution procedure in Maline and
Imidazolium based DESs: Eight solid U matrices namely UO3, Uranyl nitrate (UN)
UO2(NO3)2, UO2, U3O8, Rb2U(SO4)3, U-metal, (U, Pu)O2 and U3Si2 and one PuO2 were
dissolved in Maline (Table S2.1). The dissolution process of solid matrices of UO3, Uranyl
nitrate (UN) UO2(NO3)2, UO2, U3O8 in Maline was carried out by ultrasonication nearly 3
hours afterwards heating at 60°C about 1 to 2 hour in METALAB oven. The dissolution of
Rb2U(SO4)3, U-metal and U3Si2 matrices were also carried out by ultrasonication and heating
steps. The ultrasonication of solid matrices Rb2U(SO4)3, U-metal and U3Si2 in Maline were
performed nearly 3 to 5 hours and the heating step was done at 60°C for 2 to 3 hours. The
dissolution of PuO2 and (U, Pu)O2 matrices were carried out in the radioactive glove box
53
using IR lamp. IR heating (~100 °C) of PuO2 and (U, Pu)O2 matrices were performed in the
repetitive stepwise manner of heating and cooling such as 10 to 15 minutes heating and half
an hour cooling. The repetitive heating and cooling were done for 3 to 4 hours till the
complete dissolution of solid matrices. The number of samples taken and their corresponding
concentration was reported in Table S2.1. Weighed amounts of uranium oxides (UO3 and
UO2) were added to the DES 1 and DES 2 separately and were sonicated (~ 3-4 hours) till the
entire solid was dissolved. In the case of UO2, in addition to sonication, heating at 700C for 1
hour was also carried out for complete dissolution. The concentration of dissolved uranium
2.3 Summary of used Techniques: The brief discussion about the various techniques
transient electrochemical technique where the applied potential (E) on the electrode is
changed with time and current (i) flowing through the electrode is measured as a function of
to the concentration of the analyte. These techniques are inexpensive and strictly analytical
(i.e., can determine the low levels of metals in different matrices) and can be classified into
three categories, which are sweep techniques, pulse techniques and hydrodynamic
solution the sweep techniques are preferred while quantitative information about the analyte
can be obtain by pulse techniques. In voltammetry, the main instrument used is a potentiostat
which provides a potential to the electrochemical cell and the resulting current is then
measured. Modern potentiostat contains a built-in sweep and/or pulse generator, and those
which are interfaced to a computer to generate the desired waveform. Potentiostat use a three-
54
[75-77]
electrodes. The voltammetric techniques are based on Fick’s and Faraday’s Laws .A
simple electrode reaction Ox + ne- ↔ R proceeds through a series of steps as discussed below
2. Electron transfer between the metallic electrode surface and the analyte (Electron Transfer)
3. Movement of the product away from the electrode surface to the bulk solution
4. Coupling of different chemical reactions preceding or following the electron transfer (e.g.,
protonation, dimerization, catalytic decomposition etc.) and other surface reaction (e.g.,
The mass transport of redox species towards electrode can be taken place through any of the
following processes:
1. Migration: The transport of charged species under potential gradient due to electrostatic
force between the charged electrode and the analyte ions. The migration of analyte can be
55
2. Diffusion: The movement of anayte towards working electrode under concentration
gradient.
3. Convection: The transport of analyte due to the mechanical motion of the solution by
stirring of the solution, rotating electrode or vibration of the solution or electrode. It is usually
eliminated by keeping the experimental set up under quiet and stable condition.
analyte in the solution. The power of CV lies in its capability to quickly provide information
the potential of a stationary working electrode is scanned linearly versus time from Ei
(starting potential) to Ef (switching potential) and reversing back to Ei, using a triangular
potential-time (E-t) waveform [Fig. 2.4]. Single or multiple cycles can be used depending on
the requirement of information. The potentiostat records the current resulting from the
applied potential during the potential sweep and the associated plot of current versus
potential, thus obtained, is known as a cyclic voltammograms. The scan rate (ν) of CV is the
rate at which the potential of the WE vary with time. Figure 2.4, shows the characteristics CV
response of a reversible electron transfer process of analyte during a single potential scan. For
an analyte, the redox parameters such as anodic peak current (Ipa), cathodic peak current (Ipc),
anodic peak potential (Epa), cathodic peak potential (Epc) and half wave potential (Ep/2,
potential corresponding to half of peak current) can be attained from CV. There are mainly
two primitive types of behaviour, as discussed below, which controls the peak current of an
56
Figure 2.4. E-T curve and cyclic voltammetry plot with varying concentration profiles 1 to 7.
results in the concentration gradient between the bulk solution and the electrode. Therefore,
diffusion of analyte occurs towards the electrode. The observed current depends on the
=>
7 = !9:; <= ? (2.2)
@A
Where I is the current (A), A is the area (cm2), D is diffusion coefficient (cm2 sec-1),
(BC/Bx)x=0 is the concentration gradient (mol cm-1) at the electrode surface, C is the
concentration of the analyte (mol cm-3) and x is the distance from the electrode to the bulk
(cm). In presence of applied cathodic potential, an analyte starts to reduce and current flow.
Beyond the formal reduction potential, the surface concentration of an analyte becomes
nearly to zero and the mass transfer to the surface reaches a maximum rate (current
approaches its maxima, Ipc) and afterwards it declines due to the depletion effects
57
[(BC/Bx)x=0) decreases]. Therefore, a peaked current potential curve is obtained. Mass
F2G
E4/5 = E A (K L K , ) (2.4)
* H / IJ
(M#F)2G
E+ 1 = EA (K L K , ) (2.5)
* H / IJ
Where, k0 = standard rate constant, α = transfer coefficient (0< α< 1) and n is the number of
electrons transferred. The heterogeneous standard rate constant k0 measures the intrinsic
ability of a species (say, Ox) to exchange electrons with the electrode in order to convert to
its redox partner (say, Red). A large k0 implies that a species will convert to its redox partner
occurs on a long-time scale in case of a small k0. The current-potential relationship which is
half-reaction.
58
irreversible, and quasi-reversible depending on the electron transfer rate. The shape of CV
changes accordingly, as shown in Fig. 2.5. Compared to the rate of the mass transport, if the
rate of the electron transfer is higher, the process is defined as electrochemically reversible
(both kred and kox are large and greater than the rate constant of the mass transport). In another
condition, where the rate of the electron transfer is lower than the rate of the mass transport,
the process is defined as electrochemically irreversible. The diagnostic criteria for reversible,
Table 2.1: Diagnostic criteria for ET-processes in terms of various redox parameters
0.059
)V
XKY = ZKY3 L KY6 Z Ep shifts with ν and shift is
2.303 %\
XKY > (
1. of about 30/αn mV for a n
=
!9 tenfold increase of ν XKY i!dm"gf"f hiPℎ k
0.059
=( ) V at 25°C
n
XKY de!fPg!P hiPℎ k
(KY3 + KY6 )
K, =
2.
2
3. Ep is independent of scan Ep shifts with ν and shift is
rate (ν) of about 30/αn mV for a
tenfold increase of ν
7Y6 7Y6
=1 = 1 e!o ip q = 0.5
4. No current ratio (Ipa/Ipc)
7Y3 exists 7Y3
Otherwise
7Y6
< 1; q > 0.5
7Y3
7Y6
> 1; q < 0.5
7Y3
5. 7Y ∝ k M/ 7Y ∝ k M/ Current-function
vw
Current-function Current-function z W { |i}ℎP !eP ~"
vw vw
(xW/y ) if de!fPg!P hiPℎ k (xW/y) if de!fPg!P hiPℎ k xy
de!fPg!P hiPℎ k
59
Pulse Voltammetry: The pulse voltammetric techniques lowers the detection limits of
increasing the ratio between the faradic and non-faradic currents. The difference between the
various pulse voltametric techniques is the excitation waveform and the current sampling
regime. For speciation studies, more sensitivity pulse techniques are differential pulse
voltammetry and square wave voltammetry which are discussed in detail as below.
Differential pulse voltammetry (DPV): It is extremely useful and sensitive technique for
measuring analyte at trace level. In DPV, fixed magnitude pulses superimposed in a linear
potential ramp are applied to on working electrode and the current is sampled twice just
before the pulse application [i(t1)] and at end of the pulse [i(t2)]. The current difference [∆i =
i(t2)-i(t1)] is plotted versus the applied potential and the resulting peak shaped curve is called
differential pulse voltammograms [Fig. 2.6]. This procedure effectively reduces the
background current and thus this procedure results in a Faradaic current free of most
capacitive current. The major advantage of DPV is low capacitive current, which leads to
high sensitivity. The peak current (ip) is directly proportional to the concentration of the
Where σ = exp [(nF/RT)(∆E/2)] (∆E is the pulse amplitude). The width of the peak (at half
Figure 2.6. (a) The waveform of DPV and (b) the differential pulse voltammogram.
60
In DPV, the irreversible redox system results in lower and broader peak current with inferior
potential and is applied to the working electrode. The current is sampled twice during each
square wave cycle, first at the end of the forward pulse (at t1) and second at the end of the
reverse pulse (at t2). The difference between the two measurements is plotted versus the base
staircase potential [Fig. 2.7]. In SWV, the frequency (fsw) and scan rate (ν) is correlated to the
M
p0 = (2.8) and k= = p0 XK0 (2.9)
-w -w
where XK0 is the staircase shifts at the beginning of each cycle. The equation for actual peak
Where, ∆Ψp is the dimensionless peak current and depends on n (number of electrons), ∆Ep
Figure 2.7. (a) Potential modulation (b) Single potential cycle in square-wave voltammetry
61
The SWV has several advantages such as excellent sensitivity because the net current is
larger than both the forward or reverse components and the rejection of background currents.
The SWV is 4 and 3.3 times higher for reversible and irreversible systems, respectively, than
the analogues DPV. Major advantages of SWV are its speed and increases in the signal to-
noise ratio. It is useful the study of electrode kinetics to and determination of some analyte at
Chronoamperometry (CA): In CA, a fixed potential is applied to the working electrode and
the current produce is measure with time. It generally involves stepping of potential at the
working from a value at which no faradic reaction occurs to a final potential at which the
surface concentration of the electroactive species is effectively zero [Fig. 2.8]. The resulting
working electrode and unstirred solution are used and under these conditions the mass
transport happen mainly by diffusion, therefore, chronoamperogram shows the change in the
concentration gradient in the vicinity of the surface. With progress of time, the slope of the
concentration profile of analyte is decreased due to gradual expansion of the diffusion layer
associated with the depletion of the analyte. Consequently, the current decays with time
2G> W/y
i (P) = = EP #M/ (2.11)
W/y - W/y
CA is frequently used for measuring the diffusion coefficient of analyte, surface area of the
working electrode and mechanism of the electrode processes, especially the ECE type. In
62
Figure 2.8. (a) Potential-time waveform (b) change of concentration profile with time (c)
Chronopotentiometry (CP): In CP, the current is held constant and the potential becomes
the dependent variable, which is measured as a function of time. The steady current (i) is
applied to the working electrode through which causes analyte (Ox) to be reduced at a
constant rate. The potential of the electrode moves to values characteristic of the Ox/Red
couple and varies with time as the Ox/Red concentration ratio changes at the electrode by the
continuous flux of the electrons. This consequences in a E-T curve like the that obtained in
potentiometric titration [Fig. 2.9]. After the concentration of Ox drops to zero at the electrode
surface, the flux of Ox to the surface is insufficient to accept all of the electrons being force
across the electrode-solution interface. The potential of the electrode will then rapidly shift
toward more negative values until a new, second reduction process can start. The time after
application of the constant current when this potential transition occurs is called the transition
time (τ), which is related to the concentration and the diffusion coefficient of analyte
The shape and location of the E-t curve is governed by the reversibility, or the heterogeneous
rate constant, of the electrode reaction. The CP is useful for mechanism and heterogenous
ET-rate constant for stepwise electron transfer reaction, ET with coupled chemical reactions
63
Figure 2.9. (a) Current-time waveform (b) typical potential-time response
focused spectroscopy. It is coupled method which have the advantages of both the
electrode surface [Fig. 2.10]. Principally, any form of spectroscopy such as IR, UV-Visible,
Raman, EPR, NMR etc., can be used to follow the progress of an electrode reaction.
A light beam passes through the electrode surface to measure the absorbance changes
resulting from species produced or consumed in the electrode process. Therefore, optically
transparent electrodes (OTEs), which enable light to passed through their surface and the
64
adjacent solution, are the prerequisite in SEC experiments. OTEs may be (i) thin film (100-
quartz, or plastic substrate or (ii) fine wire metal micromesh “minigrids” containing small
(10-30 µm) holes, which can provide good optical transmission (> 50%) and electrical
conductivity. For a general redox process: O + ne- ⟷ R, the change in absorbance is related
to concentration and optical path length. The absorbance change can be described by
considering a segment of solution of thickness dx and if R is the only absorbing species, the
differential absorbance upon passage of the light is dA = εRCR(x, t)dx, where εR is the molar
absorptivity of R. The total absorbance is then : = ∫A C (, P)&, which is total amount
∞
of R and is equal to Q/nFA (Q= is the charge passed in electrolysis). Since Q is related to
The above equation implies that absorbance is linearly dependent on t1/2 (and respective A-t
curve is called as chronoabsorptometry) and useful to determine the diffusion coefficient and
εR.
Applications: SEC is very useful for (i) the elucidation of reaction mechanism, and for the
explanation of kinetic and thermodynamic parameters, (ii) measuring the formal redox
potential and number of electrons transferred during redox conversion, (iii) the mechanism
of a potentiostatic circuitry and a voltage ramp generator) and a recorder (i-E-t). Some basic
65
1. Electrochemical cells and electrodes: The electrochemical cell is generally an enclosed
beaker of 2-50 mL volume, which contains the three electrodes namely as working, reference
and auxiliary. These electrodes are immersed in the analyte solution. The electrode at which
the reaction of interest occurs is known as working electrode (WE) while the electrode that
provides a stable and reproducible potential and against which the potential of the WE is
bridge to avoid the contamination of the analyte solution. A current carrying electrode of
inert conducting material such as platinum wire or graphite rod is employed as auxiliary
electrode (AE). The relative position of these electrodes and their proper connections to the
Working Electrode: The characteristics of WE should be such that it should provide high
signal-to-noise ratio, reproducible response, chemical inertness with analyte, low cost, non-
mechanical properties. The mercury, carbon, noble metals such as platinum, silver and gold
long period of time is better for RE. The normal hydrogen electrode/Standard hydrogen
electrode (E0=0 V), saturated calomel electrode (SCE: Hg/Hg2Cl2/KCl [saturated], E0= 0.242
V vs. NHE) and silver-silver chloride electrode (Ag/AgCl/KCl [saturated], E0= 0.197 V vs.
NHE) are some commonly used aqueous based REs. For the non-aqueous media non-aqueous
Counter Electrode: The area of the CE must be larger than that of WE so that its area does
not control the limiting current. Also, CE should not get dissolved in or react with the
experimental solution and the reaction product during redox reaction at the CE should not
react with WE. Pt-wire or glassy carbon are frequently used as CEs.
66
2. Solvents and supporting electrolyte: The choice of the solvent is governed by the
solubility of the analyte and its redox activity, and by the solvent properties such as the
products, and the wider potential range without any side reactions in presence of potential. In
electrochemical experiments, the frequently used solvents are water and some non-aqueous
required to decrease the resistant of the solution, to maintain constant ionic strength,
swamping out the effect of the variable amount of naturally occurring electrolyte, and to
eliminate the electromigration effect i.e., The SEs, generally, are inorganic salts, a mineral
acid, or buffers. An excess of unreactive ionic salts (often termed as “swamping electrolyte”),
about 100 times greater than the concentration of analyte, is required for purely diffusion-
controlled transport of analyte towards electrode and to avoids its migration from bulk.
3. Inert gas purging: The electrochemical reduction of oxygen usually proceeds via two
electrode)
+ 2 + 2" # → 2 (2.15); E1/2 = -0.9 V (Versus Sat. calomel
electrode)
The large background current accruing from above redox reaction of oxygen interferes with
the reduction process of analyte and the products of oxygen reduction can also affect the
electrochemical process of interest. Therefore, for gaining information about the true redox
peaks of analyte, the oxygen removal from the solution is utmost task. Commonly, the
purging with an inert gas, such as high purity nitrogen gas, for 5-10 minutes prior to
67
4. Electrochemical analyzer: The schematic diagram of a three-electrode potentiostat is
given in Fig. 2.11. Basically, the electrochemical analyzer instrument contains two circuits as
(i) a polarizing circuit which applies potential to the cell, and (ii) a measuring circuit which
monitors the cell current. The analyzer cares the potentiostatic control of the working
electrode and minimizes errors from cell resistance (R). The applied potential (Eapp) is related
to ohmic potential drop (iR) as Eapp = EWE-ERE-iR. The ohmic potential drop results in poorly
defined voltammograms with lower current response and shifted and broadened redox peaks.
Under potentiostatic control, compensation of major fraction of the cell resistance is achieved
loops. In addition, the reference electrode is placed closer to the working electrode and is
connected to the instrument through a high resistance circuit that draws ultimately no current
from it. A current carrying auxiliary electrode is placed in the solution for completion of
current path and current flow through the solution between the WE and AE. Therefore, no
current passes through the RE and due to it specific position (as close as possible to WE), the
potential drop caused by the cell resistance (iR) is minimized. The closeness of RE with the
Weis accomplish using a Luggin probe, which is as specially designed bridge of the RE. In a
scenario, when the potential sensed by the RE is lesser than the desired value, the operational
current to voltage converter (called as “current follower”) to the WE, it if feasible to measure
the current without disturbing controlled parameters. With a ramp generator in instrument,
uncompensated solution resistance between RE and WE and due it some fraction of potential
drop called as iRu is still be included in the measured potential. This fraction is large when
resistive non-aqueous media are used. However, using appropriate positive feedback, this
68
fraction of drop can be automatically compensated from the potential signal. Bi- and multi-
Instrumentation of electrochemical methods: For uranyl with Citrate, PPA and PCs ligands,
the cyclic voltammetry (CV), differential pulse voltammetry (DPV) and square wave
systems consisting of SMDE as working electrode (sensitive for speciation analysis [79]), Ag-
AgCl/(sat. KCl) as a reference and glassy carbon as an auxiliary electrode. For Np-PPA
system, the glassy carbon electrode (GC) was used as working electrode while other
electrodes remain same as above. In case of studies in DES media, the CV, DPV, and in-situ
Before use, the GC working electrode was polished with graded 5µM alumina powder and
Figure 2.11. (a) Three electrode set-up (b) schematic diagram of a three electrode
69
High purity N2 gas was used to purge the solution for at least 10 minutes prior to each
absorption Cuvette with the Teflon top was used as a spectroelectrochemical cell. ITO plate
was used as the working electrode with Ag/AgCl (sat. KCl) as the reference electrode and Pt
elementary tool to get qualitative information about the functional group of organic moieties
and oxidation state of metal ion in solution. In addition, it is one of the direct techniques for
quantitative analysis of soluble species since the spectra obtained are the representative and
= ( ) = (2.16)
Where, A is the absorbance, I0 and I are the intensity of incident and transmitted light, ε is the
the concentration and path length), c is the molar concentration of the absorbing species and l
is the path length. The species (metal ion, ligand and complex) present in the solution absorb
the incident photon and exhibit the transition from ground state to excited state. The
wavelength at which maximum absorption occur is denoted as λmax and is used for
measurements were performed using a JASCO V530 model spectrophotometer. The V-530 is
a double beam spectrophotometer which contains the light sources as a deuterium (D2) lamp
(190 to 350 nm) for the UV region and a halogen (WI) lamp (340 to 1100 nm) for the Vis-
NIR region. It measures the absorbance of a species in the wavelengths range 190 to 1100 nm
70
with a wavelength resolution and accuracy of 1 nm and ± 0.5 nm, respectively. The half-
neutralized ligand solution is added in increments to a fixed volume (2.0 mL) of actinyl ions
solution present in quartz cuvette. The spectra are recorded for each addition of ligand
solution to the metal solution and enough time (300 s) was given for equilibration (though the
reaction was instantaneous) after each addition before recording the spectra. All the titration
data was analyzed using the Hyperquad computer programming [80] to determine the stability
2.3.3 EXAFS: The X-ray Absorption Fine-Structure (XAFS) is the modulation of the x-ray
absorption coefficient at energies near and above an x-ray absorption edge. The XAFS
spectroscopy has been widely employed for studying the local structure around a metal ion at
molecular scale. The term “XAFS” is a broad one that comprises mainly XANES (X-ray
Absorption Near Edge Structure) and EXAFS (Extended X-ray Absorption Fine Structure)
[Fig. 2.12]. The physical quantity that is measured in XAFS is the X-ray absorption
coefficient µ(E), which describes how strongly X-rays are absorbed as a function of X-ray
energy. XAFS is based on the X-ray photoelectric effect, in which an X-ray photon incident
on an atom within a sample is absorbed and liberates an electron from an inner atomic orbital
(e.g., 1s). The “photoelectron” wave scatters from the atoms around the X-ray absorbing
atom, creating interferences between the outgoing and scattered parts of the photoelectron
X-ray absorption probability, which is proportional to the X-ray absorption coefficient. The
XAS measurements have been carried out at the Energy-Scanning EXAFS beamline (BL-9)
at Indus-2 Synchrotron Source (2.5 GeV, 100 mA) at Raja Ramanna Centre for Advanced
[81]
Technology (RRCAT), Indore, India . This beamline operates in the energy range of 4
KeV to 25 KeV. The beamline optics consists of a Rh/Pt coated collimating meridional
cylindrical mirror and the collimated beam reflected by the mirror is monochromatized by a
71
Si(111) (2d=6.2709 Å) based double crystal monochromator (DCM). The second crystal of
DCM is a sagittal cylinder used for horizontal focusing while a Rh/Pt coated bendable post
mirror facing down is used for vertical focusing of the beam at the sample position. Rejection
of the higher harmonics content in the X-ray beam is performed by detuning the second
crystal of DCM. In the present case, XAS measurements have been performed in both
transmission mode and fluorescent mode. For the transmission measurement, the absorption
where, x is the thickness of the absorber. The EXAFS spectra of all samples at U L3-edge
were recorded in the energy ranges of (17050-17900) eV. In order to take care of the
defined as follows:
µ (E ) − µ0(E ) (2.17)
χ (E ) =
∆µ0(E0)
where, E0 is absorption edge energy, µ 0 ( E0 ) is the bare atom background and ∆µ 0 ( E0 ) is the
step in µ ( E ) value at the absorption edge. The energy dependent absorption coefficient
χ ( E ) has been converted to the wave number dependent absorption coefficient χ (k ) using
the relation,
2m( E − E0 )
K= (2.18)
h2
where, m is the electron mass. χ (k ) is weighted by k2 to amplify the oscillation at high k and
the χ(k)k2 functions are fourier transformed in R space to generate the χ ( R ) versus R spectra
in terms of the real distances from the center of the absorbing atom. The set of EXAFS data
[82]
analysis programme available within Demeter software package have been used for
EXAFS data analysis. This includes background reduction and Fourier transform to derive
72
[82]
the χ ( R ) versus R spectra from the absorption spectra using ATHENA software and
finally fitting of experimental data with the theoretical spectra using ARTEMIS3 software.
The data has been fitted from 1-3.2 Å in R space and the goodness of fit has been
R factor
determined by the value of the defined by:
where, χ dat and χ th refer to the experimental and theoretical χ (r ) values respectively and Im
and Re refer to the imaginary and real parts of the respective quantities.
Energy (eV)
Figure 2.12. The representative XAFS spectra showing various regions of XAFS (XANES,
2.3.4 Calorimetry: Isothermal titration calorimetry (ITC) is a technique used to study the
by changing the chemical composition of the sample by titration of a required reactant. The
heat associated with the reaction is the direct thermodynamic observable (related to both the
enthalpy and extent of reaction). The foundation of this technique is based on the
measurement of heat, q. Because of the relationship between the change in enthalpy and the
73
change in internal energy (∆E) of a system, the heat (q) absorbed or released is equal to ∆H
∆H = ∆E + P∆V (2.20)
Thermometric AB, Sweden) was used to carry out all the calorimetric experiments. It is a
twin thermopile heat conduction type calorimeter and differential power signal measured is
dynamically corrected for the thermal inertia of the system. The titration assembly consists of
4 ml reaction vessel and a reference vessel. The heat capacity of reaction vessel and reference
vessel is balanced by keeping the same volume of solutions in both sides in order to minimize
the short-term noise. The titrant is delivered in the reaction vessel through a stainless-steel
injection needle (length 1 m and internal diameter 1.5 х 10-4 m) connected to the Hamilton
syringe containing the titrant. The syringe is driven by Lund Syringe pump. The temperature
can be obtained by either titrating the protonated ligand solution with base or by titrating
deprotonated ligand anion with acid solution. In the present study, the later procedure was
employed to determine the enthalpy of formation of protonated species of ligands. All the
titrations were carried out in 1.0 M NaClO4 ionic medium at 298 K temperature. At each step
of titration, the reaction heat (QiP) solely of protonation, was calculated using the equation.
Where QiD is the heat change during dilution titration, QiT - the sum of heat changes due to
both the protonation and dilution. The heat change during the formation of protonated ligand
74
is related to ∆HP (enthalpy of protonation) and ∆Hn (enthalpy of neutralization) by following
relation
QiP = ∆HP1 (viHL – vi-1HL) + (∆HP1 + ∆HP2) (viH2L - vi-1H2L) + ∆Hn ∆vHn0 (2.24)
X
where vi - the number of moles of species ‘X’ in the cup after ith injection which were
calculated using Hyperquad Simulation and Speciation software (HySS 2006). For the
calorimetric titration of complexation studies, the metal ion solution at the desired pH and
ionic strength was taken in the reaction vessel and was titrated with the ligand solution. The
heat released due to complexation only (Qi C) at any injection i can be given by
Qi C = Qi T - Qi D - Qi P (2.25)
T D
where Qi is the heat released in titration of ligand with metal atom and Qi is the heat
released due to dilution and Qi P is the heat of protonation of the ligand. Using Hess's law of
constant heat summation, the enthalpy of formation for each complex (for example: the
(ESI) is a soft ionization technique extensively used for production of gas phase ions without
speciation (basically the stoichiometry of metal-ligand complex) of metal ions in solution [83].
The basic architecture of the ESI-MS composed of three basic components as ion source,
mass analyzer, and detector. The intact molecular ions are produced in the ionization
chamber where the ion source is kept, and then they are transferred in the mass analyzer
region via several ion optics, which are basically kept to focus the ion stream to maintain a
stable trajectory of the ions. The mass analyzer sorts and separates the ions according to their
mass to charge ratio (m/z value). The separated ions are then passed to the detector systems to
measure their concentration, and the results are displayed on a chart called a mass spectrum.
Since the ions in the gas phase are very reactive and often short lived, their formation and
75
manipulation should be conducted in high vacuum. For this reason, the ion optics, analyzer,
and also the detectors are kept at very high vacuum (typically from 10-3 torr to 10-6 torr
hypodermic needle or stainless-steel capillary. A very high voltage (2-6 kV) is applied to the
tip of the metal capillary relative to the surrounding source-sampling cone. Because of this,
the dispersion of the sample solution into an aerosol of highly charged electro-spray (ES)
droplets occurs. The ions derived by ES process are multiply charged, and the analyte
remains intact (no fragmentation). In positive ion mode the charging generally occurs via
protonation (sometimes metalation also), but in negative ion mode charging occurs via
deprotonation of the analyte. During transfer of an analyte from solution to the gas phase via
solution undergoes
the high-voltage
analyte solution is
solvent evaporation
(from the charged droplet) and droplet disintegration, resulting a very small charged droplet,
which is able to produce the charged analyte, and (c) finally a mechanism by which the gas-
Instrumentation of ESI-MS: Mass Spectrometric studies were carried out with the
76
analyzer (model micrOTOFQ-II, Bruker Daltonics GmbH). Data analysis was carried out
with Bruker Compass Data Analysis software (supplied by Bruker Daltonics GmbH). The
solution was infused into the mass spectrometric system with the help of 100 µL syringe
(Hamilton) using a syringe pump (NEMESYS, Cetrol GmbH). The capillary voltage of -4500
V and +3800 V were applied for the measurement in the positive ion mode and negative ion
mode, respectively. Nitrogen was used as both dry gas and nebulizer gas at a flow rate of 4
theory (DFT) and molecular dynamics simulation (MD) are the important methods (i) to
understand the coordination modes of ligand with actinyl ions, (ii) to get the optimized
structure with minimum energy of species with ligand in aqueous medium and DES media
(iii) to know the hydrogen bonding interaction in DES formation using HBA and HBD,
number of hydrogen bonds with dilution of DES with water and the possibility of formation
the experiment results via calculating the diffusion coefficient, bond distances, partial charges
DFT calculations for aqueous system: The structure of three CPs and their complexes with
UO22+ ion were optimized without any symmetry restriction using B3LYP functional with
triple valence plus polarization (TZVP) basis set as implemented in Turbomole package.
Relativistic effective core potential (ECP) was used for U where 60 electrons are kept in the
core of U. The energies were calculated at the same level of theory. Due to the inclusion of
non-local HF contribution in the exchange functional, the B3LYP functional works well in
predicting the energetics. The frequency calculations were carried at the same level which is
used for geometry optimization. The gas-phase free energy, ∆Gg was computed at
T=298.15K and P=1atm. The solvent effects were considered using the Conductor like
77
Screening Model (COSMO) solvation model. The default COSMO radii were used for all the
elements. The dielectric constant, ε of water was taken as 78.4. The gas-phase minimum
energy structures were used for the calculation of single-point energy in the COSMO phase.
MD and DFT in DES media: For the simulation of UO22+ ion in the solution of choline-
malonic-water, first the geometries of choline and malonic were first optimized using M06-
2X functional [84] and TZVP basis set [85] as implemented in Turbomole suite of package [86].
Then a simulation box was created with 500 choline, 500 chloride, 1000 malonic acid (MA)
(for pure DES as well water-saturated DES), 5 uranyl (UO22+) ions and 10 Cl- molecules.
Percentage of water was kept 5%, 10 and 20% by placing water molecules in the box.
(dimensions of approximately 6.5 × 6.5 × 6.5 nm3). Initial configurations of the system were
[87]
built with the PACKMOL package , ensuring a random distribution of molecules. In the
[88]
present study, all the simulations were performed using GROMACS (VERSION 5.1.2)
molecular dynamics simulation package employing all-atom optimized potentials for liquid
simulations (OPLS–AA) for bond stretching, bond angle bending and dihedral angle torsion.
For, UO22+, we have used the potential parameters from our earlier study [89]
and for water,
SPC water model was adopted. The systems were first energy minimized using the conjugate
gradient method and then the steepest descent method of minimization. All ions were
represented by charged 12-6 Lennard-Jones sites bound together by harmonic bond distance,
bond bending (angular), and dihedral potentials. Lorenz-Berthelot rules were used throughout
for the mixed Lennard-Jones terms and the long-range electrostatic interactions were
incorporated using the particle mesh Ewald techniques. The systems were simulated for 100
ps and 1ns of equilibration in NVT and NPT ensemble respectively using Noose-Hoover
thermostat and Parrinello-Rahman barostat for pressure coupling with a time step of 1fs.
Further, the data was collected after the production run of 50 ns using NPT ensemble. The
78
simulation trajectories were visualized using VMD graphical package using data of 1ns run
(i) Thermogravimetry analysis (TGA): The loss in weight as a function of temperature for
pure DES and uranium matrices dissolved DESs was recorded in thermogravimetry and
Mettler Toledo Pvt Ltd, Switzerland was used. Commonly the loss in weight occurs as a
result of the breaking of bonds and liberation of various molecules. The weaker bond breaks
at a lower temperature and stronger at a higher temperature. Prior to the actual measurements,
the instrument was calibrated for temperature and heat flow rate using high purity standards
like In, Pb and Zn. 40 mg of each sample were loaded in the sample crucibles of capacity 150
uL and run at a heating rate of 10 Kmin-1 with an argon flow of 50 ml min-1 from 25oC to
450oC. A blank run was carried out with the same experimental conditions to subtract
atmospheric and instrumental effect by keeping two empty alumina crucibles of capacity 150
uL in reference and sample holders. The melting point of pure components and DESs were
measured using DSC (DSC823e supplied by M/s Mettler Toledo Pvt Ltd, Switzerland) with a
heating rate of 5 Kmin-1 and in argon gas flow of 50 mlmin-1. The temperature and heat flow
calibration were carried out following a similar procedure as mentioned above for the TG-
DSC system.
(ii) D.C. Arc AES and EDXRF: The D.C. Arc atomic emission spectrometric (D.C.Arc
AES) technique was carried out using Spectro-Arcos Atomic Emission Spectrometer
procured from Spectro-Arcos, Germany. The array of capacitively coupled device (CCD) was
used as a detector system instead of a photomultiplier tube. The resolution of the instrument
was recommended as 0.02 nm for the wavelength range 450-800 nm. The optical design
corresponds to Paschen Runge mounting. The spectrometer has holographic grating with
79
maximum groove density 3600 grooves/nm. The focal length of the optical arrangement is
750 mm. Using this instrument, the maximum wavelength covered was 130-800 nm. In D.C.
Arc AES, standard graphite electrode, ASTM designation E-130-66 type S-2 on a type S-1
pedestal were used as the anode. ASTM designation E-130-66 type C-1 was used as a pointed
cathode. A Jordan Valley EX-3600 M EDXRF spectrometer was used for XRF analysis. The
system is equipped with a Si(Li) detector, which is having a 12.5 mm thick beryllium
window. The nEXt system software provided by the manufacturer was used for acquisition
and analysis of the spectrum. The detector used in the present case was having a resolution of
139 eV at Mn-Kα (5.9 keV). The wavelength range under investigation is 1–40 keV. 30 kV
high voltage and 200 µA emission current was utilized for data acquisition. The cross-
sectional area for the aperture of the X-ray beam at the sampling site is circular in nature with
(iii) FT-IR spectra: FT-IR spectra were recorded using the Platinum attenuated total
reflection (ATR) technique (ALPHA FTIR spectrophotometer). All spectra were obtained
using a resolution of 4 cm-1 (wavenumber) and equal measurement conditions (3900–450 cm-
1
) and 40 scans. (Scan means 16 repetitions of a single FT-IR measurement).
(iv) Karl-Fischer titrations (KF): The water content in DESs were determined by Karl-
Fischer titration using a single batch of KF reagent procured from Merck by Titrando-905
(Metrohm AG) auto-titrator. The instrument was calibrated by titrating a known weight of
water in methanol using the KF reagent and deriving a multiplying factor F with units of mg
of water /volume of KF in mL. Known weights of DESs were similarly titrated and
(v) Speciation modelling: In the present studies, speciation modelling of actinyl ions in
presence of ligands were carried out using two software namely, Hyperquad simulation and
80
speciation (HySS) software [90] and MINTEQA2 [91]. HySS is a computer program written for
the Windows operating system on personal computers which provides (a) a system for
simulating titration curves and (b) a system for providing speciation diagrams. MINTEQA2
[92]
is an equilibrium speciation model which uses the Newton-Raphson approximation
method and includes the implementation of a competitive Gaussian model for computing the
complexation of metals.
The heat change (Qir) at each addition of ligand to the metal solution is the summation of heat
due to metal-ligand complexation (Qic), ligand protonation (Qip) and the dilution (Qid)
processes. The heats due to solely complexation can be deduced by following relation.
(a) The number of protons associated with electron transfer step in both region A & B was
accounted from following equations with using their corresponding slope values.
Where p and n is the number of protons and electrons respectively involved in ET step, α is
(b) With the assumption of coupled chemical reaction, the slope of the linear portion of Ecom
Following the Laviron equation indicates the linear dependency of Ep on the logarithm of
81
.A ¦H .A
E = E A + < (α¡)¢
? log < (α¡)¢ ? + < (α¡)¢
? logν (2.30)
where A is the electrode area, F the Faraday, icom the reductive peak current, C0* the bulk
electroactive species, the potential of the working electrode at any time (t) can be
RT nFAC 0 k 0 f , h RT t (2.33)
E (t ) = log[ ]+ log[1 − ( ) 1 / 2 ]
α nF i α nF τ
Where, kof, h is the heterogeneous forward rate constant and E(t) is the potential at time
If Ef is known and Ep occur beyond Ef, a plot of lnip versus (Ep-Ef) would give a straight line
with a slope of (αnF/RT) and intercept of ln (0.227 nFAC0*k0). The value of αn and k0 were
evaluated from slope and intercept of the straight line of lnip versus (Ep-Ef) plot respectively.
Where Ep is the peak potential, Ef is the formal redox potential, k0 is the standard
electrons, F is the Faraday constant, ν is the scan rate (V/sec), R is gas constant (8.314 J K−1
(g) The linear relationship between Ep and log ν according to the following equation:
º º
°± ¹ ²³´½
® = − ¯
[. µ¶ + . ·· ¸ ¼ + . ·· < ? ] (2.35)
²³´ » °±
82
Where k0 is the standard heterogeneous rate constant and all other notations have the usual
meanings.
(h) In those cases where the calculation of Ef is not so simple, it is derived from the literature
[92]
that equation 2.36 can also be used, eliminating the Ef, for the calculation of k0.
#º/
(¾ )¿℃ = º. ºº Á «Âà L ÂÃ/ ¬ ĺ/ (2.36)
Where the notations have their usual meanings, knowing values of Ep-Ep/2 at a
particular scan rate and with the knowledge of D0, equation 2.36 can be used to
calculate k0.
(iii) Equation used in Molecular dynamics and DFT simulation: It is very important and
necessary to understand the microscopic structure of the molecular system and thus the pair
correlation functions (PCF) have been evaluated to examine the bulk structures of molecular
Here, N, r and ρ indicate the total number of molecules, distance and number density
respectively. Ni and Nk denote the number of molecule i and molecule k respectively and rik is
from Einstein’s equation using the mean square displacement (MSD) profile as:
Where r(t) and r(0) are positions of atom at any time t and at t=0 respectively. The left part of
the equation indicates the ensemble average of squared displacement i.e., MSD. The
2.6 Determination of stability constant of complex: The methods used in present studies
83
(i) Electrochemical method: The voltametric studies of stabilities of metal complexes,
complexing ligands. The shift increases with increasing ligand concentration. Out of a
number of possible methods for mathematical analysis of shift in potential, the Deford-Hume
method has been applied here owing to the possibility of successive complex formation with
increasing ligand concentration. In this case the peak potential was found to shift with
varying ligand concentration at fixed metal ion concentration or varying metal ion
concentration at fixed ligand concentration. The stability constants can be determined by the
Where, F0(X) is a polynomial function representing the sum of the βn[X]n for all complexes,
βn is the cumulative stability constant of the nth complex, [X] is the analytical or total
concentration of the added species (PPA in this case), and β0=1 for the zeroth complex. F0(X)
Where, ∆Ep = (Ep)s-(Ep)c, that is, the shift in potential; n and Ip are number of transferred
electrons and peak current, respectively. The subscripts c and s indicate complexed and free
anion respectively. The evaluation of other functions from F0(X) is described elsewhere in
detail [93] and they are mentioned briefly here for convenience:
84
At constant ionic strength and pH, the shifts in potential with increasing ligand or metal ion
concentration have been measured and used to evaluate the F0(X). From the plot of F0[X.] vs.
taking known concentration of a metal ion and a ligand, and experimentally monitoring the
optical densities of the metal, ligand & metal-ligand mixture separately. Metal-Ligand
complex stoichiometry can be obtained by the Job's method of continuous variation, method
[94-95]
of mole ratio, method of slope ratio, method of isosbestic point . The absorbance of the
solution at any point of time is an additive contribution of the absorbance of all individual
species present such as metal, ML, ML2…MLn in the solution as shown below,
Where εMLi is the molar absorption coefficient of species MLi, CMLi is the concentration of
species MLi. Each incremental addition of ligand solution to metal solution consequences in
an above equation for all the wavelengths and results in a matrix of the form A (N, λ).
Herein, N is the number of titration point and λ is the wavelength of consideration. During
titration, the solution of this matrix in combination with mass balance equations are used to
determine the molar absorption coefficients and stability constants of the species formed. The
The knowledge about the appropriate green metrics is a prerequisite for a meaningful
85
(a) E-Factor: It is the quantity of waste that is produced for a given mass of product. It also
product. Final product in this context applies to a single chemical transformation, a series of
× 100
Ñ+)/3()64 /1Õ.- +, Y4+5(3-
:Pe| Kde!e| (%) = (2.49)
Σ Ñ+)/3()64 /1Õ.- +, 4/63-62-0
(c) Mass intensity: It considers the yield, stoichiometry, the solvent, and the reagent used in
the reaction mixture, and expresses on a weight/weight basis rather than a percentage
economy (AE), yield and the stoichiometry of reactants are included. RME is the percentage
Ñ600 +, Y4+5(3-
%"gdPie! |gff "ppidi"!d (%K )(%) = Σ Ñ600 +, 4/63-62-0
× 100 (2.51)
(1) provides information about different oxidation states of metal ion in solution (ii) low
detection limit (Table 2.2) and low cost (iii) has large linear range and ability to speciate
based on complex eciatlability (iv) low energy excitation technique and species selective
rather than element selective (v) equally suitable for analyzing inorganic, organometallic and
organic compounds.
86
Table 2.2 Detection limit of various techniques for speciation of actinides
87
Chapter 3
MS studies on solutions of varying UO22+ to ligands ratios and pH were used to confirm
complex stoichiometries.
88
3.1 Introduction:
Speciation studies on Uranium have been of considerable interest for quite some time
[16-17. 19-20, 97]
with an objective to predict its distribution and fate in the environment . This
becomes important as anthropogenic activities, such as, milling and mining, may create
[20]
situations of higher concentrations of uranium in the geosphere than that naturally
abundant in soil and water (2-4 ppm and 3 ppb respectively). Its radiological and
toxicological effects have been the subject of interest to environmentalists. Hence, knowledge
of the speciation of uranium and its complexes is one of the interesting subjects, with regard
to understanding its migration behavior in an aquifer [16-17, 20, 27] and to evaluate its impact on
the living organisms. Uranium mobility and transportation in the environment is strongly
dependent on its oxidation states, and hence, there has been renewed interest in redox
chemistry of uranium and its complexes in the recent past. Uranium is a redox-sensitive
[20, 98]
element and its redox chemistry is exceptionally rich and diverse due to readily
accessible oxidation states ranging from III to VI and the sensitivity of the redox potential to
[98]
the coordination environment around the metal ion . Investigation of the redox properties
of uranium in presence of chelating agents [20, 26, 29] is important from the point of view of its
bioavailability, and toxicity, which are remarkably dependent on the oxidation states and
forming strong water-soluble complexes with both U(VI) and U(IV) over a wide range of pH
[99]
. It is a versatile chelating agent, and is a common constituent of domestic, industrial as
well as nuclear wastes. With proper steric arrangement it forms small chelate rings in the
equatorial plane of the uranyl ion giving highly stable uranyl chelates. It exhibits relatively
consistent removal efficiency which has been extensively used for decontaminating the
89
contaminated sites in nuclear facilities. Uranyl ion complexation with citrate in
availability for microbial and phytocoenosis groups [100-101]. Citric acid is preferred over other
organic ligands for removal of uranium due to the formation of bio- or photo-degradable
[100]
complexes of uranium thereby controlling its mobility in the environment . Citric acid
also important in chelation therapy towards the development of formulations for detoxication
and excretion of uranium [20]. The pka values of citric acid are given in scheme 3.1.
attention because of their potential applications in ion exchange, catalysis, sorption, and
properties, which are unique, and promise more efficient separation processes for waste
clean-up and environmental restoration [103-104]. Organic and inorganic phosphorus containing
ligands which are pervasive in soil systems are known to form strong complexes with
phospholipids, phosphonic acids, phosphoric mono- and diesters in soils rich in organic
[105]
matter . Though relatively low content of P (800−1500 mg/kg) is found in humic
substances, organic compounds containing P=O and P−OH functionalities show excellent
[106]
complexing tendencies towards metal centre and are presumed to play a remarkable role
in binding of uranyl ion by HS. The complexation tendency of phenylphosphonic acid (PPA)
[104]
for uranyl ion is stronger than that reported for the isostructural carboxylic acid . Higher
stability of the phosphonate complexes of the uranyl ion with regard to the corresponding
carboxylate is due to differences in net electrostatic charge and basicity between the
phosphonate and the carboxylate ligands [107] as well as an entropic effect of the phosphonate
90
unit having more potentially binding oxygen atoms than that in carboxylates [108]. Further, the
cycle of uranium in soils, its dissemination in surface waters and the mobility in phosphate-
rich soil systems [103-104, 109]. The pka values of PPA are given in scheme 3.1.
Scheme 3.1. The Pka values of (a) Citrate (b) Phenylphosphonic acid (PPA)
Earlier investigation on the Uranyl Citrate complexes have been reported in the literature
very mobile, sensitive to the solution conditions, and arduous to accurately characterize,
resulting in the lack of comprehensive understanding of its behavior despite decades long
research as many discrepancies among published data exist and knowledge gaps remain. Of
presence of ligand. Moreover, despite the extensive spectroscopic studies mentioned above,
91
no systematic study has been carried out on the speciation of UO22+ in Citrate medium
considering its redox behavior. With PPA ligand, the redox speciation of uranyl is
electrochemical method) to present the systematic study on the speciation and redox
behaviour of UO22+ with varying physicochemical conditions in aqueous Citrate and PPA
medium. The present study signify its important for providing a new direction in analytical
methodology to gain about the metrical knowledge viz. speciation, in presence of chelating
agents at low concentration [Herein 10-5 M] of metal ion. The present work involves study of
redox speciation of Uranyl in presence of Citrate and PPA by Voltammetry, with emphasis
metal ligand ratios. Attempts have been made to obtain the stoichiometry of the complexes,
3.3. Results and Discussions: The experimental results and related descriptions are provided
uranyl ion concentration and pH in presence of citric acid is given in Fig. 3.1 a-b. The
plots show a significant shift in peak potential of U(VI) in presence of Citrate with
observations, the speciation diagram for Uranyl Citrate system at lower (10-5M) and
higher (10-3M) concentration of UO22+ was calculated using the geochemical code,
92
MinteqA2 [91]. The speciation diagrams [Fig. S3.1a-b] show that monomeric [UO2Cit]-
species dominates the speciation over other species in the pH range of 2.5 to 8 at lower
UO22+ concentration (10-5M) while in the same pH region the dimeric [(UO2)2Cit2]2−
species dominates the speciation over other species at higher UO22+ concentration (10-3
were carried out at both UO22+ concentrations. The mass spectra acquired for the
Uranyl Citrate system as a function of pH and UO22+ and Citrate concentration are
given in Fig. S3.2 a-e. The intense peak at m/z value 459 [Fig. S3.2 a-b], corresponds
to monomeric [UO2Cit]- species. In gas phase the dimer can be singly or doubly
charged anionic or neutral depending on the protonation state of the two free carboxyl
groups. The intense peak at m/z= 919 in Fig. S3.2 c-d corresponding to
UO22+ concentration. Based on the speciation diagram and the ESI-MS spectra, the
cyclic voltammograms [Fig. 3.1a] obtained at lower UO22+ concentration [10-5M] can
be explained as due to reduction of monomeric [UO2Cit]- species while the same [Fig.
dimeric [(UO2)2Cit2]2− species. These conclusions are in agreement with the report by
29
Ohyoshi et.al. (1975) that at lower UO22+ concentration (≤10-5M) polymerization of
uranyl ion is diminished and monomeric species, namely, [UO2Cit]− predominates the
speciation. The conclusion about the dominance of dimeric species [(UO2)2Cit2]2− over
monomeric species at higher UO22+ concentration (≥10-3 M), is also in agreement with
[110-113]
the literature reports based on spectroscopic techniques . It has been observed
93
[(UO2)3Cit3]3− and (UO2)3Cit2. However, under these conditions, reductions of
In order to obtain better understanding about redox behavior of both the species, the
reduction pattern was investigated with varying pH. As shown in Fig. 3.1a-b both the
which the two peaks get merged and finally single reduction peak appears at pH ≥ 6.
This indicates the change in the redox behavior of the monomeric and dimeric Uranyl
Citrate complexes with pH. Referring to the speciation diagram in Fig. S3.1 a-b it can
above 5 and the monomeric and dimeric species dominate the speciation upto pH 8.
The ESI-MS spectra [Fig. S3.2 a-b] show that the intense peak at m/z= 459 remains
constant at both pH values (4 and 6.5). Further, the intense peak at m/z= 919 also
m/z values with pH. Thus, the reduction patterns obtained at pH 4 and 6.5 in Fig. 3.1
(a) and (b) can be attributed to reduction of monomeric and dimeric species
respectively. The different redox behavior of same species with varying pH could be
of uranium in aqueous solutions has been widely studied earlier [116-120] and it has been
reported that in acidic medium the first reduction peak is due to reduction of U(VI) to
[119]
U(V). Suzuki et.al reported slow dispropotionation of U(V) in acidic pH (2-5),
thereby explaining slight stability of U(V) in this pH region, while at pH >5, the
U(V). Since Citrate can stabilize U(V), the second reduction peak (peak II) can be
94
obtained, which can be explained in terms of the fast disproportionation of U(V), and
The redox potential (Eh) vs. pH plots for monomer and dimer obtained from CV are
given in Fig. 3.2 (a) & (b), respectively. Change in the peak potential with varying pH
indicates the involvement of protons during the reduction process. For both the species
plot I is the variation of first reduction peak [U(VI) to U(V)] potential upto pH 5.5 and
the two electrons reduction peak [U(VI) to U(IV)] potential at pH ≥ 6, while plot II
shows the variation of second reduction peak [U(V) to U(IV)] potential with pH up to
5. Plot I show two linear regions with discontinuity at pH 5.5, while plot II shows a
single linear region upto pH 5. The discontinuity at pH 5.5 in peak I observed in both
the species could be due to change in redox mechanism as discussed earlier. The slope
of linear portion of Eh-pH plot is related to the number of protons involved during
reduction and is given by the equation 2.27 for irreversible reduction at pH ≤ 5 and
equation 2.28 for reversible at pH ≥ 6. The observed value of slope of linear region A
and B and corresponding number of protons calculated for both species are given in
Table 3.1, which shows that for both species, no proton is involved during first
reduction step [U(VI) to U(V)] at pH ≤ 5, while one proton is involved during second
reduction step [U(V) to U(IV)] upto pH 5. The number of protons during [U(VI) to
For both the species at pH 4, the plot of peak potential (Ep) vs. logarithm of scan
rate (logν) for first reduction peak [U(VI) to U(V)] are given in Fig. 3.3 (a) & (b). The
Ep values were found to vary linearly with logν in case of both the species, which,
along with the absence of oxidation peak in reverse direction of applied potential [Fig.
3.1a-b], confirms the irreversibility of reduction process. The value of αn can be easily
calculated from the slope of Ep vs. logν plot. Assuming the first step as
95
electrochemical reduction coupled with a chemical reaction (described later), the slope
of linear portion of Ep vs. logν plot can be given by equation 2.29. The value of αn
calculated from the observed slopes for monomer (-0.0568) and dimer (-0.0588) using
equation 2.29 are 0.519 and 0.501, respectively. Assuming α as 0.5, the value of n was
calculated as 1 for both the species, thus corroborating the first reduction peak as due
to U(VI) to U(V) conversion. For monomeric and dimeric species, the ip vs √ν plot at
pH 4 (first reduction peak) and pH 6.5 [Fig. 3.3(b)] depicts the linear relationship
between ip and √ν, indicating diffusion control reduction process of both the species at
both pH values. The diffusion coefficient (D0) of [UO2Cit]- and [(UO2)2Cit2]2− were
and 2.32 (used at pH 6.5) [121-122] and are given in Table 3.2.
The Square wave voltammograms recorded at three pH values, namely, 1.5, 2.5 and
3.5 for UO22+ concentration of 10-3 M and Uranyl to Citrate ratio of 1:1 are shown in
Fig. 3.1c. The figures indicate a single reduction peak at pH 1.5 and three reduction
peaks at pH 2.5 and 3.5. Peak I (Ep= -0.152) is due to the reduction of free UO22+
absence of any ligand. However, it is not clear if the reduction peaks II and III
obtained at pH 2.5 and 3.5 are due to reduction of monomeric or dimeric species. The
Eh-pH diagrams for peak II and III, shown in Fig. 3.2 (c) and (d), respectively, help in
values. The Eh-pH plot for peak II shows two linear regions A (Slope = -0.046) and B
(Slope = -0.048) with discontinuity at pH 3. Both the linear regions have nearly the
same slope which is comparable with the slope of Eh-pH plot for the first reduction
peak of monomer and dimer species at pH ≤ 5 (Table 3.1). Therefore, peak II at pH 2.5
and 3.5 can be attributed to U(VI) to U(V) reduction with the discontinuity due to
96
change in nature of species above and below pH 3. The speciation diagram [Fig.
S3.1c] at same UO22+ concentration and Uranyl to Citrate ratio implies that the three
range 1.5 to 3.5. The Fig. S3.1c also shows that below pH 3 monomeric species
dominates the speciation while at pH>3, the speciation is dominated by the dimeric
species. The ESI-MS spectrum in Fig. S3.2e indicates the existence of both the
monomer (m/z = 459) and dimer (m/z = 919). Therefore, it can be inferred that the
reduction of monomer and dimer dominate the reduction process at pH <3 and pH>3
respectively. The Eh-pH plot for peak III is shown in Fig. 3.2d. The slope of linear
region C was found to be -0.110 and nearly equal to that of second reduction peak of
of U(V) to U(IV).
explore the kinetic behavior of reduction for Uranyl Citrate species. Herein the
reduction follows the ECE mechanism for both Uranyl Citrate species. In cyclic
voltammograms over a wide range of potential scan rates, no anodic currents are
observed in case of monomer and dimer at pH ≤ 5, indicating both the reduction steps
to be irreversible and coupled with the chemical process (an ICI mechanism). The
consideration of data in Fig. 3.3c-d with increasing scan rates and small currents
permits some qualitative features of the coupled chemical reaction to be deduced. The
effect of varying scan rates on the value of peak current functions (i/√ν) for the first
reduction peak for both monomer and dimer have been summarized in Fig. 3.3 (c),
97
decrease in the same is observed indicating complex kinetics for both the Uranyl
Citrate species. The decrease in current function with increasing scan rate differentiate
the observed reduction mechanism of the present case from that involving only multi
step charge transfer and no chemical reaction between charge transfer steps, as for the
[75-76]
latter case the peak current functions (i/√ν) should be invariant with scan rate .
Thus, present observation indicates the existence of chemical reaction between charge
transfer steps.
For first reduction peak [U(VI) to U(V)], the plots of chronopotentiometric constant
(iτ1/2, where τ is the transition time for reduction wave) vs. applied currents for a series
of pH values for both 1:1 and 2:2 Uranyl Citrate species are summarized in Fig. 3.3 (d)
and the amount of applied current (i). iτ1/2, for a diffusion-controlled electron transfer
Thus, iτ1/2 should be independent of the magnitude of applied current. However, Fig.
3.3d shows that iτ1/2 is not constant and varies inversely with applied current and
approaches a constant value at high applied current. This suggests the involvement of
complex kinetics in the first reduction step with iτ1/2 varying with actual concentration
[124-125]
of the electroactive species . The potential of the working electrode at any time
by equation 2.33. A plot of E(t) vs. log [l - (t /τ)½] will be a straight line with slope of
59/αn mV and an intercept that gives the value of kof, h. The graph for both 1:1 and 2:2
Uranyl Citrate species have been plotted in Fig. 3.4 (a), respectively. Curve A in Fig.
3.4 (a) has been obtained from data at low current densities [in the kinetic region
98
shown in [Fig. 3.3d] and reveals appreciable deviation from theory as t approaches τ.
While, curve B in Fig. 3.4 (a) results from data in the region of constant iτ1/2 [Fig.
3.3d, current-limiting value] and deviates from linearity for both large and small
values of t. The value of αn and kof, h calculated from slope and intercept of the linear
portion of the curves are given in Table 3.3. The values of E(t=0) have been evaluated
by extrapolation of the straight-line portion of the E(t) vs log [1 –(t/τ)1/2] to t=0. The
graph of E(t=0) vs. log i for monomer species is shown in Fig. 3.4d. The plots
indicates that at high applied current, where iτ1/2 is constant, E(t = 0) is a linear
function of log i but at the lower current (kinetic) region a systematic deviation from
linearity takes place. The same behavior is also observed in E(t=0) vs. log i plot for
dimeric species. The E(t=0) shifts cathodically with increasing current, which also
process is ruled out as the voltammetric wave for a disproportionation process would
exhibit an anodic shift of about 30/αn mV per tenfold increase in the potential scan
51
rate whereas in the present case, exactly opposite effect, that is, a cathodic shift has
been observed [Fig. 3.3 (a)] with nearly 30/αn mV per tenfold increase in scan rate,
[126]
which is characteristic of an ECE mechanism . The chronopotentiometric constant
(iτ1/2) and peak current function (i/√ν) at constant current density and constant scan
rate respectively, initially increase up to pH 3.5 and subsequently decrease for both the
Uranyl Citrate species [Fig. 3.4b-c]. These two quantities (iτ1/2 & i/√ν) are function of
concentration of electroactive species, and hence the initial increase in iτ1/2 and i/√ν up
and dimeric species, which follow the same pattern in the speciation diagram [Fig.
S3.1a-b]. However, the same argument does not hold for the pH region 3.5-5, wherein
the concentrations of the momoner and dimer remain constant, but the functions, iτ1/2
99
and i/√ν have a decreasing trend, indicating a decrease in the amount of the electro-
active species generated by the chemical reaction with decreasing hydrogen ion
concentration. This suggests that the chemical reaction step appears to involve
protonation of the product of the first reduction step [123-124] as given below,
3.3.3. Reduction Mechanism: The citrate ligand does not produce any
the issue due to the free citrate ligand on the overall electrochemical reduction
fixed UO22+ (10-5 M) was also examined on the redox behaviour of UO22+-Citrate
complex. Shift in peak potential and change in redox behaviour were not observed
with changing concentration of Citrate from 10-5 M to 5x10-3 M. This confirms that
Citrate. As evident from Fig. 3.1 (a) & (b), no change in redox pattern was observed
with change in the UO22+ concentration indicating the conversion of dimeric species,
firstly into monomer in the electrode surface region followed by the same reduction
At pH ≤ 5.5
[(UO2)2Cit2]2− 2[UVIO2Cit]−
At pH ≥ 6
100
[(UO2)2Cit2]2− 2[UVIO2Cit]−
understand the redox behaviour of other Uranyl complexes. Herein, mono oxo Uranyl
and 6.5, where no kinetic complication is observed and the system follows 1e- and 2e-
reduction, respectively. When the same experiment is carried out at pH 3.1, the
mechanism. The observed current is intermediate between 1e- and 2e- reduction, since
reaction of [UVO2Cit]2- to generate [HUVOCit]+. Fig. 3.5b displays the data in terms
of normalized i vs. t-1/2. The shape of the ICI curve can be understood from i vs. t-1/2
plot in terms of the time taken for the chemical reaction involving [UVO2Cit]2-. At
shorter times (larger t-1/2) the ICI curve approaches the 1e- Cottrell plot [i= nFAC0*
√D0 ⁄√πt] due to insufficient time for formation of [HUVOCit]+ from [UVO2Cit]2-.
longer times (small t-1/2), there is adequate time for the formation of [HUVOCit]+
101
3.3.5. In Situ Spectroelectrochemical study: With a view to corroborate the ICI
an optically transparent cell. The absorption spectra were measured at each applied
potential in the range from -0.01 to -1V with varying time of electrolysis. The results
are shown in Fig. 3.6. The spectra recorded after 25 min. electrolysis with varying
potential from -0.01 to -1V are shown in Fig. 3.6a. As seen from Fig. 3.6b no clear
protonation of the species produced during the first reduction step. Furthermore,
during electrolysis the absorption bands at 487, 504, 560 and 663 nm, which are
consistent with absorption peaks of U(IV) in Citrate, become intense with time and
with decrease in applied potential as shown in Fig. 3.6a. The reduction of U(VI) is also
evident from the decrease in intensity of absorption bands of U(VI) in Citrate at 422,
436 and 447 nm. Above observations suggest that the reduction of U(VI)Citrate
species proceeds through two steps reduction process: the first reduction step
protonation of the species generated during first reduction step. This chemical reaction
102
through U(V), one might expect the existence of the U(V) species in the Citrate
medium. However, the literature reports on the signature of U(V) complexes are
show the presence of new U(V) species in aqueous Citrate medium, namely,
[UVO2Cit]2-. We observed the appearance of the absorption bands at 732, 743, 774,
785, 798 and 982.8 nm associated with the reduction of U(VI)Citrate to U(V)Citrate as
shown in Fig. 3.6c-d. The intensity of absorption bands increases with time from 0 to
25 minute at fixed potential (-0.01V) and the trend continues with decrease in potential
from -0.01V to -0.3V up to 1 minute after which the intensity starts to decrease with
further increase in reduction time as well as decrease in applied potential from -0.3V
to -1V [Fig. 3.6c-d]. This type of variation in intensity of absorption bands with
reduction time (at fixed -0.01 V) and varying (-0.01V to -0.3 V up to 1 minute)
reduction time from 1 to 35 minute at -0.3V the second reduction step initiates
faster with decrease in potential from -0.3 to -1V, resulting in decrease in the intensity
0.3 to -1V. Literature reports on absorption bands of U(V) in visible and near-infrared
[132-133]
regions, show peaks at 765 and 980 nm in K2CO3 aqueous solution .
Furthermore, the characteristic absorption bands of U(VI) [422, 436 and 447 nm] and
U(IV) species [487, 500, 560 and 663 nm] occur at wavelengths well below 700 nm.
103
Therefore, the absorption bands at 732, 743, 774, 785, 798 and 982.8 nm, obtained in
the present study have been attributed to [UVO2Cit]2-, which arise due to f-f transitions
3.3.6. Determination of stability constant: Voltammetric studies does not show any
significant shift in peak potential with varying ligand concentration from 10-5 to 10-2
Citrate from UO22+ with no change in the nature of the species with increasing ligand
concentration was varied from 10-6 to 10-3 M at pH 4. The plot of first reduction
[U(VI)-U(V)] peak potential (Ep) vs. logCM is given in Fig. 3.7a, which shows two
linear regions A and B having different slopes. This result further confirms the
the reduction of dimeric species. The stability constant of [UO2Cit]− was determined
[93]
following the Deford-Hume formalism . At constant ionic strength I=0.1 M and pH
4 the shifts in potential with increasing UO22+ concentration have been measured. The
plot of F0[UO22+ Conc.] vs. [UO22+ Conc.] (equation 2.41) is given in Fig. 3.7b, which
represents a straight-line having slope = 6.3x106. The logβ value of monomer was
31
deduced as 6.8±0.11, which is in excellent agreement with the literature data . The
first reduction step in the redox process is expressed by the following equation:
104
[UVIO2Cit]− + e- [UVO2Cit]2− (3.2)
RT β V (3.3)
E f = E0 + ln
αnF β VI
Where, E0 is the standard redox potential, while βV and βVI represent the stability
Literature reports [107-108, 134] indicate that UO22+ in PPA medium exists mainly in three
forms exist in solutions at pH 3-7 while a 1:2 UO2(HPhPO3)2 complex is formed additionally
concentration of UO22+ and 10−3 M concentration of PPA at pH 4 and pH 5 are given in Fig.
3.8 (i-ii). The figures show a significant shift in peak potential of U(VI) in presence of PPA
with respect to that in absence of PPA recorded at same metal ion concentration, indicating
complexation of U(VI) by PPA. At pH 4 the difference between potential of peak I and that
of U(VI) in absence of PPA is observed to be (∆Ep)I = 0.102 V while the same for peak II is
to corroborate this observation, the speciation diagram for U(VI)-PPA system was calculated
[91]
using the geochemical code, MinteqA2, . The speciation diagram (Fig. S3.3) shows that
1:1 U(VI)-PPA complex dominates the speciation over other possible species in the pH range
of 2 to 7, thereby indicating that the CV in Fig. 3.8 is due to reduction of UO2PhPO3 (1:1)
complex.
105
3.3.7. Redox behavior with pH: Variation in redox behavior of UO2PhPO3 (1:1) species has
been observed with change in pH values. As shown in Fig. 3.8 (iii-iv) in 3D plots similar
changes have been observed in CV and DPV with variation in pH values. Upto pH 4 two
reduction peaks are obtained with intensity of second reduction peak being higher than first
one. The electrochemistry of uranium has widely been studied in aqueous solutions earlier
[115-120]
and it was reported that in acidic medium the first reduction peak is due to reduction
of U(VI) to U(V). Accordingly, in Fig. 3.8(iii-iv), peaks I and II can be assigned to reduction
of U(VI) to U(V) and that of U(V) to U(IV), respectively. Higher intensity of peak II than
compared to that of U(V). However, at pH 4.5 a drastic change in redox behavior has been
observed with the intensity of peak I being higher than that of peak II, indicating that at pH >
4 PPA has a tendency to stabilize U(V) oxidation state over U(IV). The disproportionation
[135] [58, 135]
of U(V) to U(IV) and U(VI) described by the following reactions may be
study indicates that in the presence of PPA, the slow disproportionation of U(V) continues as
[119]
expected upto pH 4 but beyond pH 4 the fast disproportionation, as proposed earlier ,
does not occur in this case. This suggests that in presence of PPA at pH > 4 U(V) will be
stable with respect to disproportionation. Therefore, two reduction peaks upto pH 4 could be
106
explained in terms of slow disproportionation and less stability of U(V), while the single
and more stability of U(V). This further explains why the peak intensity of U(VI) to U(V)
reduction become higher while that of U(V) to U(IV) lower at pH 4.5 and this reduction peak
[U(V) to U(IV)] totally vanishes after pH 4.5 as can be seen in Fig. 3.8 (ii). Redox behavior
of 1:1 species at three pH conditions viz., acidic, neutral, and alkaline are given in 3D plots of
CV and DPV [Fig. 3.8 (v-vi)]. Two reduction peaks at acidic condition and single reduction
peak both at neutral and alkaline condition are evident from figure.
3.3.8. The Eh-pH plot: The Eh-pH (Eh = redox potential which is function of H+ ion) plot
provides information about the existence of species in particular pH region and number of
protons involved during reduction. The Eh-pH plots for UO22+ (10-5 M) in presence of PPA
(5×10-3 M) for both the reduction peaks are given in Fig. 3.9(i) and (ii). The plot for peak I
(Fig. 3.9 (i)) shows two linear regions with discontinuity at pH 4.5, while peak II (Fig. 3.9(ii))
shows single linear region upto pH 4. The speciation plot shows the existence of single
dominant species UO2PhPO3 (1:1) in the pH region from 2 to 7, indicating no change in the
nature of species at pH 4.5. Hence the discontinuity at pH 4.5 could be due to change in
redox mechanism. The slope of linear portion of Eh-pH plot is related to the number of
where, P = no. of protons involved in reduction. The slope of linear portion of region A for
peak I was found to be -0.048 V using equation 2.27, while the number of protons involved
during reduction was calculated as 0.4, which can be neglected due to its insignificant value.
This suggests that no proton is involved during this reduction step. The corresponding values
107
of the slope and number of protons obtained from the linear portion of region B for peaks I
and II were found to be -0.074 V and 1.2 and -0.062 V and 1.0, respectively.
3.3.9. The Ep vs. log ν plot (determination of heterogeneous kinetic parameter in acidic
values were found to vary linearly (R2 = 0.99) with log ν for reduction peak I (region A) at
pH 3 as shown in Fig. 3.10(c). Further, the absence of oxidation peak in reverse direction
(Fig. 3.8(i)) confirms irreversibility of reduction peak I upto pH 4. The value of αn can be
easily calculated from the slope of Ep vs. log ν. For the data plotted in Fig. 3.10(c), the slope
is 0.096, which gives αn as 0.61, which is close to 0.5, the value of α for irreversible
reduction and confirms the single electron (n = 1.2) transferred in the reduction process. The
value of k0 can be determined from the intercept of the same plot (-0.372) provided E0 is
known. The value of E0 in equation 2.30 obtained from the intercept of Ep vs. ν curve, was
peak in reverse direction of applied potential at pH 5. The ratio of reduction to oxidation peak
current was calculated to be (1.3), which is close to unity. Furthermore, Fig. 3.10(d) shows
that the reduction potential is not affected significantly on varying the scan rate. These results
[98]
are characteristics of reversible electron transfer process . For a reversible reduction
process the difference between the reduction peak potential and its half value (∆Ep) is related
A.AØÙ
KY L KY× = 2
(3.10)
At 50 mV the Ep − Ep/2 was obtained as 0.047 V, which gives n as 1.21, that is, close to 1.
voltammograms for U(VI)-PPA system at two pH values 3 and 5 and at varying scan rates are
shown in Fig. 3.10 (a) and (b). In both the figures the peak current increases with scan rate
108
for peak I, confirming diffusion control reduction process. The diffusion coefficient of
UO2PhPO3 was evaluated at both pH values by using the Randles-Sevcik equation 2.31 (used
at pH 3) and 2.32 (used at pH 5) [121-122]. The value of D0 calculated at pH 5 from the slope
(7.30 × 10-8, R2 = 0.997) of linear plot of ip vs. ν1/2 (Fig. 3.10f) is 6.76 × 10-5 cm2 s-1 while the
same calculated at pH 3 from the slope (4.86 × 10-8, R2 = 0.996) of linear plot of ip vs. ν1/2
(Fig. 3.10e) is 5.03 × 10-5 cm2 s-1. Both the calculated values of D0 are in close agreement.
The heterogeneous electron transfer rate constant “ks” was calculated by using the relation
developed by Klingler and Kochi (equation 3.11), which is based on peak separation between
FH2GÜ M/ F y 2G
E0 = 2.18 Û Ý "Þ ÛL «KY6 L KY3 ¬Ý (3.11)
ª ª
Taking α as 0.5, the value of heterogeneous electron transfer rate constant “ks” was
calculated at 50 mVs-1 using D0 = 6.76 × 10-5 cm2 s-1 and found to be ks = 0.304 × 10-3 cm s-1.
3.3.13. Reduction mechanism: Based on finding of the results, the reduction mechanism of
subsequently to [UIVO2HPhPO3]−, the structure of the complex does not change as it is only
electron transfer process. However, the second reduction step is followed by the chemical
3.3.14. ESI-MS measurements: ESI-MS is used here to corroborate the number and
stoichiometries of the UO22+-PPA complexes in solution and to confirm the results from
cyclic voltammetry. Mass spectra acquired for the UO22+-PPA system are shown in Fig.
109
S3.4a, b and m/z values with corresponding species [134] are given in Table 3.4. As clear from
Table 3.4 and Fig. S3.4a, the intense peaks at m/z values 427, 445 and 463 are corresponding
to mainly UO2PhPO3 (1:1) species with its proton and one and two water adduct. Therefore,
validating the existence of UO2PhPO3 (1:1) as major species over the others in solution. This
confirms that finding of reduction peaks of UO22+ in presence of PPA in CV are because of
strength I = 0.1 M and pH 5 the shifts in potential with increasing PPA concentration have
been measured and used to evaluate the F0(PPA) using equation 2.41. Plot of the F0(PPA)
versus PPA concentration is given in Fig. 3.9(iii). The plot being a straight line (R2 = 0.992)
provides the slope = 9.54 × 106. Thus, log β value of UO2PhPO3 was found to be (6.98 ±
3.4. Conclusions
varying conditions of pH, Uranyl to ligands ratio and UO22+ concentration has been carried
out. The cathodic shift in reduction potential of UO22+ with the addition of Citrate, at both the
higher (10-3M) and lower (10-5M) metal ion concentration indicated the formation of Uranyl
Citrate complexes. At higher metal ion concentration (≥3x10-4 M) the dimeric species
monomeric species [UO2Cit]− dominates the speciation. Based on results, it has been
observed that the speciation of UO22+ in Citrate is independent of Citrate concentration and
depends only on UO22+ concentration. ESI-MS results show definitive evidence for existence
through their m/z values at 459 and 919 for monomer and dimer respectively. The
110
corroborated the existence of chemical reaction as the protonation of the reduced species
produced after first reduction step. Presence of U(V) species in Citrate medium [UVO2Cit]2-
with its characteristic absorption bands in visible and near-infrared regions at 732, 743, 774,
785, 798 and 982.8 nm were observed. Cyclic voltametry studies of U(VI)-PPA system
revealed negative shift in electrode potential with respect to UO22+, which confirms the
formation of UO2PhPO3. PPA can stabilize U(IV) upto pH 4 while above it (pH > 4) peaks
are observed owing to U(VI) to U(V) followed by U(V) to U(IV) reduction due to
reduction peak due to U(VI) to U(V) reduction occurs due to absence of disproportionation of
U(V). The m/z values at 427, 445 and 463 validate presence of UO2PhPO3 species. It may be
concluded that the identification of species UO2PhPO3 and its stability constants obtained
here, offer a broad speciation model for the U(VI)-phenylphosphonate system. This will be
useful for appraisal of the role of aromatic phosphorus functional groups in binding of uranyl
by HS.
111
Figures and Tables
(i) U-Citrate
(a) (b)
(c) (d)
Figure 3.1. Cyclic voltammetry curves: (a) [UO22+]= 10-5 M, (i) in absence of Citrate
at pH 4 (ii) in presence of [Citrate]= 5x10-3 M at pH 4 (iii) same as (ii) at pH 6.5; (b)
[UO22+]= 10-3 M, (i) in absence of Citrate at pH 4 (ii) in presence of [Citrate]= 5x10-2
M at pH 4 (iii) same as (ii) at pH 6.5 (c) Square wave Voltammetry curves with
varying pH, Condition: [UO22+]= 10-3 M, [Citrate]= 1x10-3 M (d) Variation of cyclic
voltammograms with scan rates
112
(a) (b)
(c) (d)
Figure 3.2. Eh-pH plots, Condition: (a) [UO22+]= 10-5 M, [Citrate]= 5x10-3 M and I=
0.1 M NaClO4 (b)[ UO22+] = 10-3 M , [Citrate]= 5x10-2 M and I= 0.1 M NaClO4; (c)
for Peak II (d) for Peak III, both (c) & (d) are at [UO22+]= 10-3 M, [Citrate]= 1x10-3 M
and 0.1 M NaClO4.
113
(a) (b)
(c)
(d)
Figure 3.3. (a) E vs. logν plot for peak I and (b) i vs. √ν plot (c) is the variation of peak
current functions (i/√ν) with scan rates at pH 4 while variation of
chronopotentiometric constant (iτ1/2 ) with i for several values of pH is in (d),
Conditions: (a) to (d) [UO22+]= 10-5 M, [Citrate]= 5x10-3 M and I= 0.1 M NaClO4.
114
Figure 3.4. (a) E(V) vs. log [1-(t/τ)1/2] plots at pH 4 (b) Variation of chronopotentiometric
constant (iτ1/2) with pH at fixed i = 10x10-8 A. (c) Variation of Peak Current function (i/√ν)
with pH at 50 mV/sec scan rate (d) Variation of E(t=0) with log i at pH 5.
115
Figure 3.5. Plot of i(A) vs. t(sec) with varying pH is in (a) and (b) is the plot of i(A) vs. t-1/2.
Condition: (a) and (b) [UO22+]= 1x10-5 M, [Citrate]= 5x10-3 M and I= 0.1 M NaClO4.
Figure 3.6. (a) & (b) are the absorption spectra with varying potential from -0.01 V to
-1 V after every 25 min. (c) and (d) are the absorption spectra measured with varying
both potential (from -0.1 V to -1 V) and time of pentavalent uranium. The appearance
of absorption bands are at 732, 743, 756, 774, 785, 798 and 982.8 nm respectively.
Condition: (a) to (d) [UO22+]= 4x10-3 M, [Citrate]= 4x10-3 M and I= 0.1 M NaClO4.
116
Figure 3.7. The Variation of Ep (a) with [UO22+] concentration at fixed [Citrate]= 1x10-2 M,
I= 0.1 M NaClO4 and pH 4; (b) The variation of F0[UO22+ Conc.] with [UO22+ Conc.] at pH 4.
Table 3.1. The observed slopes and calculated value of number of protons from Eh-pH
plots
117
Table 3.2. The observed slope and calculated value of D0 from i vs. √ν plot
Table 3.3. The value of αn and k0f,h for [UO2Cit]−and [(UO2)2Cit2]2−at lower and higher
current densities.
Higher 0.13 0.45 -0.15 8.69x10-6 Higher 0.13 0.45 -0.14 1.13x10-5
Current Current
(9x10-9) (7x10-7)
118
(ii) U-PPA
Figure 3.8. (i) and (ii) are the CV plots (a) UO22+ at 10-5M in absence of PPA and (b) in
presence of 5x10-3M PPA conc. (i) at pH 4 (ii) at pH 5. The 3D plots of (iii) CV at pH 4 and
4.5 (iv) DPV at pH 4 and pH 4.5. The 3D plots of (v) CV and (vi) DPV at pH 4,7 and 9.
119
(i) (ii)
(iii)
120
Figure 3.10. The i(A) vs E(V) plot with increasing scan rate from 50 to 100mV/s (a) at pH 3
; (b) at pH 5. The E vs logν plot for peak I (c) at pH 3 and (d) at pH 5. The i(A) vs √ν plot (e)
at pH 3; (f) at pH 5.
121
Table 3.4. The species with their m/z values
122
Chapter 4
AIM: Probing the systematic knowledge of neptunium (Np), such as, speciation and
aspect of its solution chemistry. The present chapter aimed at investigating the redox and
123
4.1. Introduction:
Neptunium (Np), is one of the artificial elements produced in significant quantities in
nuclear reactors by α decay of 241Am and (n, 2n) reaction of 238U followed by β decay of 237U
[137] [137-139]
. Neptunium is key constituent in spent nuclear fuel, nuclear waste and is
[140]
considered to be most problematic actinide . It remains one of the least understood major
transuranium elements though several decades have passed since the first experimental
isolation of Np (1940) [141]. It is an α-emitter with a long half-life [e.g. t1/2 (237Np) = 2.14x106
years], having high solubility under environmentally relevant conditions. Thus, the long lived
237
Np builds up in spent nuclear fuel and high-level liquid waste (HLW) with time and is one
of the important radionuclide of interest to consider for the transportation and migration in
geosphere. Np can exist in variety of possible oxidation states from III to VII in aqueous
solution and the relative stabilities of the same are strongly affected by pH and the presence
[142]
of complexing ligands . Np in its higher valence states viz. NpO2+ & NpO22+ are highly
soluble among all its oxidation states, which in turn makes these two oxidation states of Np
more mobile in geosphere. The reduction of these oxidation states to lower oxidation states
leads to its precipitation and immobilization, thus the redox speciation of Np has phenomenal
[143]
effect on its release, solubility and migration behaviour . Neptunium speciation in
aqueous media is mainly dominated by the pentavalent cation, NpO2+, under a wide range of
This redox mobility makes Np speciation highly dependent on the aqueous solution
composition and the formation of coordination complexes that disturb the Np redox
equilibrium. Further, for redox-sensitive radionuclides like Np, speciation is very complex
dependence on solution conditions and ability to coordinate with a wide range of ligands.
124
Thus, precise identification of speciation is tough, resulting in knowledge gap due to lack of
[144]
comprehensive understanding of its behaviour . The reliable spectroscopic signatures,
potentials in presence of chelating agents, are of primary concern to generate input database
for the accurate prediction of its transportation and migration in geochemical models.
Therefore, there has been a renewed charm on understanding the redox speciation of Np
[26]
species in presence of chelating agents . In this context, it is quite interesting to ascertain
significant role in actinide solution chemistry to gain knowledge about transuranic ions
are frequently used in actinide isolation and waste remediation, endure some significant
solubility and resistance for degradation [106]. The complexation tendency of Phosphonic acid
towards the metal ions is stronger than the isostructural carboxylic acid [134]. One of the most
carboxylic acid therefore metal ion can more effectively compete with H+ for truly anionic
ligand binding sites in acidic solutions [147]. Therefore, recently simple phosphonic acid based
chelating agents emerge as tenable option as sequestering agent for f-element. This ligand
assimilating doubly ionizable phosphonate groups (-PO3H2), possess many qualities which
are unique chemically and assurance for more efficient separation processes related to waste
[103-104]
remediation and environmental restoration . Furthermore, these powerful f-element
complexants can be easily destroyed at the end of their useful life by raising the temperature
or adding mild oxidizing agents, which mitigates the possibility of ligand promoted actinide
[147]
migration . High fraction of phosphorus (P) exists in organic forms as derivatives of
125
[148-149]
phosphonic acid in soils rich in organic matter . The P content in humic substances
(HS; Humic substances are naturally occurring polydisperse and heterogeneous mixtures of
organic molecules that have a high affinity for metal ions) is 800-1500 mg/kg. Therefore, the
PPA with P=O and P-OH functionalities can be acknowledged as primitive ligand for
aromatic phosphorus functional groups in humic acid. Unique features of Phosphonic acids
boost it as versatile ligands in the recent past with increasing attention due to their potential
The complexation studies on Np have been reported with inorganic ligands (CO32-,
carboxylic acids, amides, and hydroxamic acids [151-155]. The crystal studies of Np with mono
[156]
and di phosphonate [157] are also available in literature. However, there is no report in the
present work, the electrochemical and absorption spectroscopy methodology are employed to
gain insight into the speciation and coordination chemistry of Np as a function of oxidation
states and to benchmark the reduction potential and spectral characteristic of corresponding
Np species. Quantum chemical calculations (density functional theory) are carried out to
understand the possible coordination modes of Np with PPA in different oxidation states at
4.3.1. Redox behviour with pH: The cyclic and differential pulse voltammograms of Np
recorded in 0.1 M NaClO4 (pH = 2.5) using a GC electrode in absence and presence of PPA
ligand are given in Fig. 4.1. In cyclic voltammograms the two clear reduction peaks in which
first is reversible and second irreversible are obtained in both cases. The electrochemistry of
[145-146, 150, 158-160]
Np in aqueous solution has been studied earlier and reported that the
126
reversible peak is due to reduction of Np(VI) to Np(V). Therefore, in the both case the first
peaks, which are reversible, observed at 0.248 V [in absence of PPA] and 0.023 V [in
presence of PPA ligand] on the cathodic scan and their corresponding oxidation peaks
obtained at 0.392V [in absence of PPA ligand] and 0.207 V [in presence of PPA ligand] have
Further, to obtain the number of electron transfer during first reduction peak in case of cyclic
voltammograms related to Np-PPA complex the following equation 4.3 [75-76] has been used.
At 50 mV the Ep−Ep/2 was obtained as 0.070 V, which gives n as 0.81, that is, close to 1.
However, there is ambiguity in assigning the second reduction peaks in both cases. Earlier
studies explained that it may be either Np(V) to Np(IV) or Np(V) to Np(III). Atshusi et al. [8]
study on Np. In the present study, in both cyclic voltammograms, we obtained the second
reduction peaks at -0.417 V (in absence of PPA) and -0.565 V (in presence of PPA) at more
negative potential with irreversibility. This reveals that it is kinetically hindered, arduous to
reduce and involves slow electrode reaction, indicating that the second reduction peaks are
associated with reduction of NpO2+ which requires the breakage of the actinyl bonds and
release of two axial oxygen atoms of NpO2+. Hence the second reduction peaks are due to
[146]
reduction of Np(V) to Np(IV) or Np(III). Further, as explained in the previous studies
that the electrode potential of redox couple of Np(V)/Np(IV) and Np(IV)/Np(III) (only
separated by~ 0.2 V) are close hence there is possibility of overlap of following redox
reactions which results in directly two electrons transfer during reduction [146]:
127
This two electron transfer reaction can be corroborated by appearance of larger current
amplitude of second cathodic wave in cyclic voltammogram of Fig. 4.1. Therefore, the
second reduction peak obtained in both Np metal ion and its complex with PPA ligand can be
attributed to two electron reduction process namely NpO2+ to Np3+. This has also been
complex using Laviron equation 2.30. The cyclic voltammetry data analysis results related to
reversible reduction peak (Peak first) are given in Table S4.1. The table shows that ipc/ipa is
nearly fixed at 0.8 with no significant change with varying scan rates from 0.02 to 0.09 V/sec.
Further, the difference in cathodic and anodic peak potentials (∆Ep) also does not show any
measurable variation with increase in scan rates. These observations authenticate the nearly
reversibility of first reduction peak. In presence of PPA ligand a significant cathodic shift in
both peak potentials are observed in cyclic voltammogram recorded at same metal ion
of Np viz. Np(VI) and Np(V). The shift in the potential is related to the complexation
tendency of ligand. The cathodic shift (∆Ep) for peak I was obtained to be ∆EpI = 0.225 V
while for peak II it was ∆EpII = 0.148 V, corroborating strong complexation of Np(VI) by
PPA than that of Np(V). The region of existence of Np complexes with both oxidation state
viz. Np(V) & Np(VI) with varying pH are presented in the redox potential (Eh) vs. pH plots
in Fig. 4.2 (a) and (b). The linear behaviour from pH 1.75 to 3.5 as indicated by region A in
Fig. 4.2 (a) implies the region of occurrence of Np(VI)-PPA complex as a function of pH.
Below pH 1.75, there is no complexation of Np(VI) by PPA ligand while above pH 3.5 the
Np(VI)-PPA complex is unstable. The similar linear variation [region B] is also obtained in
extended upto pH 5.5 above which it does not exist in solution. The variable occurrence of
complex of Np(VI) and Np(V) with pH may be due to instability of oxidation states with pH.
128
Change in reduction potential of the complex with varying pH indicates involvement of
protons during its reduction process. The number of protons involved during reduction is
related to the slope of linear segment of Eh-pH plot by the equation 2.28 for reversible
reduction [used for Peak I] and equation 2.27 for irreversible reduction [used for peak II].
The observed value of slope of linear region A and B and corresponding number of protons
calculated using equations 2.28 and 2.27 for both complex are given in Table 4.1, which
Studies on the heterogeneous electron transfer kinetic parameters (k0, α, and D0) of
behaviour in redox environment and to elucidate the redox based remediation of soil [98]. The
of corresponding oxidation peak in cyclic voltammogram as evident from Fig. 4.1. This
[161]
irreversibility is further established by using Laviron equation , where the peak potential
(Ep) for an irreversible electrode process is expressed by equation 2.30. The Ep vs. log ν plot
grants about the irreversibility of reduction process and number of electrons involved during
reduction process and is used to find out the kinetic parameters (k0 and α). As evident from
Fig. 4.3(b), The Ep values are observed to vary linearly (R2 = 0.99) with log ν for reduction of
Np(V)-PPA complex [peak II], which confirms the irreversibility of reduction process for the
complex. The value of αn and k0 are calculated from the slope (-0.054) and intercept (-0.639)
of Ep vs. log ν plot and given in Table S4.2. The value of αn calculated from the observed
slopes is 1.09. Assuming α as 0.5, the value of n was calculated as ~2 for Np(V)-PPA
reduction process, thus corroborating the two-electron transfer reaction during its reduction
[peak II]. The heterogeneous electron transfer rate constant k0 for reduction of Np(VI)-PPA
129
[peak I] is calculated by following equation 4.6 on the basis of cathodic and anodic peak
separation [136]:
In this equation, the notations D0, Epa and Epc stands for diffusion coefficient, anodic peak
potential and cathodic peak potential, respectively. Taking α as 0.5, the value of
mVs−1 using D0 = 5.05x10-5 cm2 s−1 [reported in Table 4.2] and found to be k0 =1.73x10-3 cm
s−1.
states viz. Np(VI) and Np(V), the cyclic voltammetry measurements are carried out at
different scan rates varying from 20 to 90 mV/sec. The data are given in Fig. 4.3 (a). For both
reduction peaks [I & II] the peak current increases with scan rates. The ip vs. ν1/2 plot for the
complex Np(VI)-PPA and Np(V)-PPA are shown in Fig. 4.3 (c) and (d) respectively. The
plots (b) and (c) depict the linear relationship between ip and ν1/2, indicating diffusion control
reduction process of both complexes. The diffusion coefficient (D0) of Np(VI)-PPA and
Np(V)-PPA complexes were evaluated using the Randles-Sevcik equations viz., 2.31 (used
for peak II) and 2.32 (used for peak I) [121-122] and are given in Table 4.2.
4.3.4. Complexation mechanism: The complexation of Np(H2O)5+ with PhPO3H2 (PPA) can
be schematically represented as shown in Fig. 4.4. The mechanism involves the dissociation
of a proton from the PPA group and association of monoprotonated PPA with Np(H2O)5
130
which on further step forms complex with completely deprotonated PPA. In similar way the
The hydrated NpO2+ cation shows absorption band at 980 nm (ε = 395 M-1cm-1) due
to f-f transition which is widely used to determine the concentration of Np(V) in solution as
well as in the studies on Np(V) complexation with various ligands. The intensity as well as
position of absorption band can be used to characterize the Np(V) complexes and the
variation of absorption spectra of NpO2+ on addition of ligand can be used to determine the
stability constants of the various species formed. The spectrophotometric titrations showed a
decrease in absorbance with appearance of a single well defined isobestic point [Fig. 4.5(a-
b)] on addition of ligand. The presence of single isobestic point indicates the formation of
only one species on addition of ligand to metal ion solution. The data analysis of
spectrophotometric titration by Hyperquad 2006 [Fig. 4.5(c)] also corroborated the presence
of only 1:1 complex NpO2(PPA)- viz. [NpO2(PhPO3)]- while the Np(V) complexes with
protonated ligand were found to be absent. The determined stability constant and molar
absorbance for the Np(V)-PPA complex along with some other NpO2+ carboxylate and
dicarboxylate complexes that could resemble in mode of binding to PPA are given in Table
4.3. The f-f transitions are influenced by the ligand field around the neptunium atom which is
dominated by strong axial oxygen atoms and weaker equatorial water molecules. On binding
with the complexing ligand, the increase in ligand field at equatorial position leads to further
splitting of energy levels. The increase in ligand field results in decrease in the energy gap
between excited state 3H4 and the ground state 1G4, thereby showing the Nephelauxetic effect,
that is, red shift in the λmax on complexation is proportional to ligand field around NpO2+.
The λmax of NpO2(PPA)- shifted from 980 nm (for pure NpO2+) to 988.2 nm. The change in
λmax is higher than those of NpO2+ monocarboxylate complexes and simple dicarboxylate
131
complexes (ex: NpO2(mal)- complex) but very close to that for 1:1complex of NpO2+ with
oxydiacetate ligand (oda) wherein NpO2+ can form a 1:1 tridentate complex with oda. As the
interaction of Np(VI) with hard donor ligands are mainly electrostatic in nature, the log β of
the complex should increase with the basicity of the ligand (Linear Free Energy
Relationship). Therefore, for Np(V) complexes with hard donor atom, the ∆λmax should also
[162]
increase with the log β. Yang et al. demonstrated the linear variation of log β with ∆λmax
of Np(V) complexes with carboxylates. Fig. 4.6 shows that the ∆ λmax and log β value
The interaction of PPA with NpO2+ and NpO22+ were studied using DFT calculations.
The geometries of all the species involved in the complexation process were optimized and
interaction energy (∆E) for the complexation process was calculated. For instance, the ∆E of
reaction
àÞ ( )Ø + áℎá# ↔ [àÞ ( ) (áℎá )]
#
+ 2 (4.9)
is calculated as,
Phenylphosphonic acid (PPA) can bind with NpO2+ and NpO22+ through either deprotonation
of either one or both the hydroxyl group. In order to understand the most favourable path, the
geometries of all the possible complexes were optimized, which are given in Fig. 4.7 along
with calculated electronic charges on atoms. The energies (Ex) of all the species and
interaction energies of Np(V) and Np(VI) in aqueous solutions with both deprotonated and
NpO2(H2O)52+ (Fig. 4.7) shows that the oxygen atoms of five water molecules are located in a
plane and are nearly equidistance from Np atom (average Np-OH2O = 2.447 Å). The
132
calculated bond distances for NpO2(H2O)52+ are in good agreement with the experimentally
[163]
observed bond distances (Table 4.5) . The good agreement between theoretically
calculated with experimentally obtained bond distances validates the computation method
(Table 4.5) used in the present work. The smaller bond distances in NpO2(H2O)52+ compared
difference in the Zeff of NpO22+ and NpO2+ is larger than that for hydrated forms,
NpO2(H2O)52+ and NpO2(H2O)5+. This can be attributed to the closer approach of oxygen
atoms of water molecules and axial oxygen to Np which leads its higher charge neutralization
in NpO2(H2O)52+.
4.3.6.2 Complexation of Np(V) with PPA: The negative value of formation of completely
[PhPO3H]- and PPA2- = [PhPO3]2- respectively), ∆E values [Table 4.4], shows its higher
negative charge on PPA2- results in higher electrostatic interaction between Np and ligand
donor atoms (Fig. 4.7) which in turn results in shorter average Np-OL bond distance (2.40 Å)
in NpO2PPA- compared to that in NpO2PPAH (2.57 Å). The closer approach of deprotonated
ligand binding atoms to Np atom reduces its interaction with the coordinated water molecules
and thereby increases the average Np-OH2O (2.85 Å) compared to that in monoprotonated
species (2.57) (Table 4.5). The higher stability of NpO2PPA- is in agreement with the
4.3.6.3 Complexation of Np(VI) with PPA: Np(VI) was found to form stronger complexes
with both PPA2- and PPAH- than Np(V) which can be attributed to higher Zeff on Np atom.
This was also evident by cyclic voltammetry results where more cathodic shift was observed
in Np(VI)-PPA complex than Np(V)-PPA. The higher stability of Np(VI) complexes is also
133
reflected in the geometrical parameters of these complexes. The Np-OL and Np-Oax bond
distances in Np(VI) complexes are found to be shorter than that in Np(V) complexes. Among
deprotonated and protonated complexes, Np(VI) found to form more stable complex with
deprotonated species of both Np(V) and Np(VI), third water molecule moves away from first
coordination shell of NpO2+ and NpO22+. The strong interaction of PPA2- to Np atom leads to
reduction in its affinity to water molecule, thereby favouring the less coordinated structure.
complex. The peak potential was observed to shift with varying PPA concentration at fixed
F0(X)=antilog{[0.434αnF/RT][∆Ep]+[log(Ip)s/(Ip)c]} (4.11)
Where, ∆Ep = (Ep)s-(Ep)c, that is, the shift in potential; n= number of transferred
electrons; Ip indicates the peak current; c and s represent the complexed and free anion
respectively. The shifts in potential with increasing PPA concentration have been measured
and used to evaluate the F0[PPA] using equation 4.11 at constant ionic strength I =0.1 M. The
F0[PPA] vs. PPA concentration plot has shown in Fig. 4.8. This plot being a straight line (R2
=0.992) provides the slope (β) =6.450x103. Thus log β value of [NpVO2PhPO3]- was found to
3.84).
4.4. Conclusions
134
A systematic study is carried out to investigate the redox speciation of Np, which is in
cathodic shift in both reduction potentials of Np in presence of PPA ligand and the
bathochromic shift in λmax are clear gesture for formation of Np-PPA complexes. The
presences of new solution species of Np in its two different oxidation states (V, VI) with the
[PhPO3]2- viz. Np(V)-PhPO3 and Np(VI)-PhPO3 are identified. This study affirms the
presence of Np(VI)/Np(V) and Np(V)/Np(III) redox couples over the other probabilities
under present experimental conditions. The speciation model presented here, recommends
binding of humic and fulvic acids with Np in soil and therefore is extremely significant for
redox based studies on migration, remediation strategies of soils contaminated with Np and
135
Figures and Tables
(a) (b)
-5
Figure 4.1. (a) Cyclic voltammetry curves: [Np]=10 M and I= 0.1 M NaClO4, Black
voltammogram is in absence of phenylphosphonic acid at pH 2.5 while Red voltammogram is
in presence of [PPA]= 1x10-3 M at same pH. (b) Differential pulse voltammetry curves: at
same solution conditions as described in above.
Region A
(a) (b)
Figure 3. Eh-pH plots, (a) For reduction peak I and (b) For reduction peak II; Condition:
[Np]=10-5 M, [PPA]= 1x10-3 M and I= 0.1 M NaClO4. Region B
(a) (b)
Figure 4.2. Eh-pH plots, (a) For reduction peak I and (b) For reduction peak II; Condition:
[Np]=10-5 M, [PPA]= 1x10-3 M and I= 0.1 M NaClO4.
136
Figure 4.3. (a) Variation of cyclic voltammograms with scan rates from 20 mV/sec to
90mV/sec. (b) E vs. logν plot for Np(V)-Phenylphosphonate (Reduction peak II); Condition:
at [Np]=10-5 M, [PPA]= 1x10-3 M and I= 0.1 M NaClO4 (c) and (d) are i vs. √ν plot for
Np(VI)-Phenylphosphonate (reduction peak I) and Np(V)-Phenylphosphonate (reduction
peak II); Condition: (a)-(c) at [Np]=10-5 M, [PPA]= 1x10-3 M and I= 0.1 M NaClO4.
137
Figure 4.4. Schematic representation for the complex formation of Np(H2O)5 by PhPO3H2.
Figure 4.5. The absorbance spectra (left) and the molar absorbance (right) of the NpO2+ -
PPA complexation. ([NpO2+] = 4.23 X 10-5 M with 0.5 M PPA at pH 2)
138
Figure 4.6. Correlation between the stability constant (log β) and the magnitude of
wavelength shift (∆λmax) where, ∆λmax = λ(NpO2L)-λ(NpO2+).
Figure 4.7. Optimised structure of Np(V) and Np(VI) complexes with PPAH- and PPA2- (A)
Np(V) with PPAH-- NpO2PPAH (B) Np(V) with PPA2- -NpO2PPA- (C) Np(VI) with PPAH--
NpO2PPAH+ (D) Np(VI) with PPA2- - NpO2PPA (Red – Oxygen, Blue – Neptunium, yellow
– Phosphorous, Gray – Carbon, white – Hydrogen).
139
Figure 4.8. The variation of F0[PPA Conc.] with [PPA Conc.].
Table 4.1. The observed slopes and calculated value of number of protons from Eh-pH plots.
Reduction Peaks Slope No. of Protons
Np(VI)-Phenylphosphonate (Peak I) -0.053 1
Np(V)-Phenylphosphonate (Peak II) -0.047 1
Table 4.2. The observed slope and calculated value of D0 from i vs. √ν plot.
140
Table 4.3: The stability constant, λmax, ∆λmax and molar absorbance of the 1:1 complex of
NpO2+ with PPA along with the available literature data on 1:1 NpO2+ complexes of some
dicarboxylate (acronym full forms: mal- malonic acid, oda- oxydiacetic acid, ida-
iminodiacetic acid, ppa- phenyl phosphonic acid)
Table 4.4. Calculated energies of all the species (EX) and interaction energies of Np(V) and
Np(VI) complexes (∆E)
141
Table 4.5. Bond distances (Å) in all the Np species involved in complexation process
Bond distance
Species Bond distance (Å) Bond distance (Å) (Å)
Np-OH2 Np-PPA O=Np=Oax
NpO2+ 1.734, 1.734
1.82, 1.81
NpO2PPAH
2.60, 2.53, 2.59 (2.571) 2.44, 2.69 (2.5665) (1.817)
NpO2PPA- 2.62, 2.52, 3.42 (2.85) 2.40, 2.40(2.40) 1.83, 1.85 (1.84)
2+
NpO2
2.446, 2.445, 2.447, 1.751, 1.751
NpO2(H2O)52+
2.449, 2.449 (1.751)
NpO2(H2O)5(ClO4)2
Experimental 2.42 1.74
NpO2PPA 2.55, 2.41, 3.56 (2.84) 2.24,2.25(2.24) 1.79, 1.78 (1.78)
142
Chapter 5
AIM: In this chapter, the comprehensive speciation (redox and molecular) of uranium in its
two oxidation states namely, uranyl(VI) and uranyl(V) were investigated with three
143
5.1 Introduction
such as carboxyl and phosphonic, in PC molecule (Scheme 5.1), yields a novel chelator with
high complexing strength for metal ion through its multiple coordination sites and allows to
probe the speciation behaviour of uranyl ions in the presence of both phosphonate and
[104-106]
carboxylate groups . Its phosphonic group comprising two OH and one P=O moieties
[42-44]
has a high affinity towards actinides due to their dibasic nature . It has unique chemical
properties such as high aqueous solubility and ionization at physiological pH, greater steric
flexibility, and compelling complexing propensity even in acidic solutions at pH < 2 [104, 106,
134, 165]
. Therefore, the PC chelator would be an excellent uranophile due to the formation of
aqueous soluble complexes with UO22+ over a wide pH range and could be applicable as a
tenable option for in vivo and in vitro sequestering agent for UO22+. PC chelators are released
[166-167]
into the hydrosphere from a variety of industrial and consumer activities . Both
of glyphosate and natural organic matter, i.e., humic substances, which are environmentally
[134, 168-170]
persistent into the litho- and hydro-sphere . Furthermore, both carboxyl and
phosphonic functional groups are constituents of the cellular wall of microbe, i.e., bacteria
that are capable of surviving in highly harsh habitats like uranium waste piles [171]. Therefore,
model system for understanding the interaction of uranium (U) with glyphosate, humic
[171]
substances and the multifunctionality of bacterial surfaces . Since PC facilitating
adsorption and precipitation of metal ions [172], thus, this kind of chelator would be a suitable
144
Scheme 5.1. Structure of Phosphonocarboxylates.
In literature, previous studies are limited only to the crystal structures for
redox characteristics, and coordination modes on the interaction of UO22+ with the considered
phosphonocarboxylates, namely, PFA, PAA, and 3-PPA ligands. The aqueous speciation of
uranyl(V) with PCs is not reported in literature. The redox energetics, Eh-pH characteristic,
(VI)-PC complexes were obtained from cyclic voltammetry. The stoichiometry and
thermodynamics of interaction of uranyl(VI) (logβ, ∆G, ∆H and ∆S) with PCs have been
analysis of uranyl (VI)-PC complexes are investigated by EXAFS and DFT calculations. The
redox stability of uranyl(V) with PC in the aqueous solution was also explored with varying
pH and the stability and structure of the uranyl(V)-PC complexes were reported. The
The accurate prediction of the speciation and the basic chemistry behind the aquatic
and the protonation thermodynamic data are a prerequisite to obtain the complexation
thermodynamics data. The determined protonation constants along with the other
thermodynamic parameters are shown in Table 5.1. From, the table it could be observed that
the protonation constant values for the first protonation is around 1-2, second protonation at
3-5 and third protonation at above 7. The determined enthalpies of protonation along with
free energies calculated from log KP values and entropies of protonation for all the three
Phosphonocarboxylates are given in Table 5.1. From the table, it can be observed that the
contribution from T∆S is predominant than that of ∆H values. This indicates that the
nature of the participating chelator and the geometry arrived by its coordination on the
equatorial plane of uranyl(VI) (i.e. UO22+), has effects on the intensities and positions of
[177-178]
electronic transitions of uranyl ion . The present studies involve absorption
measurements in the titration mode, where the chelator solutions are added in increments to a
fixed volume of uranyl(VI) solution, in the range of 380-480 nm arising from vibronically
perturbed electronic transitions [177-178]. The absorption spectra profile for the complexation of
UO22+ with 3-PPA and its deconvoluted spectra showing the individual absorption spectra of
146
free uranyl(VI) and the UO22+-3-PPA complex are given in Fig. 5.1(i-ii), respectively. Similar
kind of absorption profiles and deconvolution were also observed in the case of UO22+-PFA
and UO22+-PAA complexation. The deconvolution of absorption spectra titration data was
done by HypSpec software to determine the maximum wavelength of absorption (λmax) and
the corresponding molar absorptivity (ε) at λmax of each species formed in all the three
complexation systems. All these systems showed a red shift in absorption spectra on the
addition of PCs with an increase in the absorption intensities too. The λmax with
corresponding ε values for all the species determined in the present studies are shown in
Table S5.1. HypSpec software was used to determine the stability and the species formed by
uranyl(VI)-PCs complexation. Various chemical models like (i) different combinations of the
formation of 1:1 and 1:2 uranyl(VI)-PC complexes; (ii) different protonated form of
complexes such as MLH, MLH2 and MLH3 kind of species; (iii) hydrolytic species such as
complexes of the form MLx(OH)n (x/n=0,1, 2,….); (v) the polymeric species of the form
MxLy(OH)n are tried as different models to fit the spectrophotometric data. The model
including the formation of metal complexes with monoprotonated PC chelators of the form
MLH was found to be in convergence criteria [Fig. 5.1(iii)]. The physical evidence of
speciation obtained by fitting absorbance data was further validated by ESI-MS experiments.
Fig. S5.1(i-iii) shows the ESI-MS spectra of UO22+ with the PCs and corresponding
assignments of species in terms of their respective m/z values are reported in Table 5.2. The
m/z values in the table specify the formation of MLH type of species and hence authenticate
constants are presented in Table 5.3, which indicates an interesting observation related to the
higher stability of uranyl(VI)-PCs complexes in the aqueous phase and suggests that the
mode of coordination in these complexes to be tridentate or higher. But, the literature reports
147
states that the linear UO22+ geometry restricts the participation of all the three oxygens (two
from the phosphonate group and one from the carboxylate group), which is attributed to the
electrostatic repulsion between the axial oxygen of UO22+ with the second oxygen donor of
[108]
phosphonate group . The exceptionally enhanced stability of the uranyl(VI)-PCs
between the -OH of monoprotonated phosphonate group and the axial UO22+ oxygen atom on
the complex formation (as observed in DFT calculations, described later). The log KML
values for all three UO22+-PCs are found to be nearly the same. The ring size in combination
with steric strain results in the observed trend of complex stabilities. Calorimetry titration was
employed to determine the complexation enthalpies for the interaction of uranyl(VI) with the
three PCs. The determined complexation enthalpies (equation 2.26) with all other solution
thermodynamic parameters are embedded in Table 5.3. From the table, it can be observed
that the complexation process is majorly entropy driven with very small values of
complexation enthalpies relative to T∆S values. All the complexation enthalpies are
exothermic and follow the trend of [UO2(PFAH)(H2O)3] < [UO2(PAAH) (H2O)3] < [UO2(3-
PPAH) (H2O)3]. Due to higher stability of the [UO2(PFAH)(H2O)3] complex owing to the
5.3.2.2. Redox Speciation (Redox energetic and kinetic, Pourbaix diagram, and
speciation, the solution redox behaviour of uranyl(VI) with the varying architecture of PCs
(viz. PFA, PAA & 3-PPA) were probed by cyclic Voltammetry (CV). The CV measurements
were performed to understand the influence that chelator electronic and structural changes
exert over the redox response of UO22+ due to complexation. Representative CV curves [(i)-
(v)], corresponding to UO22+ species and UO22+ in the presence of three PCs chelators,
148
obtained in the potential sweep region from 0 to -1 V, are shown in Fig. 5.2. The CV (i)
obtained for UO22+, specifies the reversible reduction wave at -0.176 V and corresponding
[116-118]
oxidation at -0.112 V, is attributed to U(VI/V) activity . Importantly, alteration in
redox potential and redox wave characteristics of UO22+ with PCs were observed in CV
curves (ii-iv). This is a clear indication of the interaction and complexation of all three PCs
with uranyl(VI) and the significant effect of equatorially attached PCs on the redox behaviour
of uranyl(VI) species. The cathodic shift in reduction wave potential was found in complexes
of UO22+ with PFA, PAA and 3-PPA chelators. The absence of oxidative wave in CV curves
(ii-iv) during anodic potential sweep [from -1 to 0 V] indicates the irreversibility of reduction
wave in complexes due to slow redox kinetics relative to that of UO22+ species in solution.
The measured peak potential (Epcom), formal potential (Efcom), the relative shift in potential
(∆Epcom) and visualized characteristic of reduction wave of UO22+ complexes with three PCs
are summarized in Table 5.4. The number of electrons associated with the reduction wave of
complexes was computed and their corresponding values have been mentioned in Table 5.4.
As obvious from the table, the number of electron is one hence the irreversible reduction
wave in the complexes is assigned to the U(VI)-U(V) process. The significant cathodic shift
in reduction wave potential of complexes than the UO22+ is consistent with the more electron-
cathodic shift in reductive wave potential of UO22+-PC complexes was found to be larger than
that of other known values of UO22+-complexes (U(VI)/U(V) redox couple) with other
Pourbiax diagrams (Eh-pH plot) of uranyl(VI) with three PCs are constructed by
experimental results [Fig. 5.3(i)]. To get an insight into the thermodynamically stable region
and to estimate the numbers of protons associated with electron transfer (ET) step of U
149
species as a function of physicochemical parameter viz. pH, it is significantly imperative to
report the electrochemical phase diagrams (Eh-pH plot). The Eh-pH plots of all three
uranyl(VI) complexes show the discontinuity in linearity in the pH region 1.5 to 8, resulting
in two linear regions A and B with different slopes [Table 5.5]. The number of protons
associated with the ET step in both region A and B was calculated from their respective slope
values (equation 2.27) and are given in Table 5.5. The number of protons in region A in all
three complexes was observed to be ~ 1. This shows the one electron and one proton
PCET (Proton coupled electron transfer) step. The speciation plots [described later] of
complexes confirm the dominancy of single complex species up to pH ≤ 6.5. At pH ≥ 6.5, the
uranyl(VI) hydroxyl species leads over the uranyl(VI)-PC complexes. However, the
uranyl(VI) hydroxyl species remain soluble in the presence of PCs in the aqueous solution up
to pH 8, as no precipitation was noticed during the experiments. Further, the observed slope
of complexes in region B is lower than that of region A [Table 5.5] and electrochemical
the change in speciation and the invariant nature of the reductive wave concerning ET in two
different regions A and B. The invariant nature of reductive wave of complexes concerning
ET in both pH regions A and B suggests that three PCs deliver sufficient stability to
nature of uranyl(V) in acidic pH (2-5) in solution, hence elucidating its slight stability in this
pH region while beyond this pH the disproportionation was rapid resulting in its
150
disproportionation was observed in near neutral (pH>5) and in the acidic region (pH <4.5) in
[180] [148]
the presence of citrate and phenylphosphonate respectively. However, with the
present PCs, an interesting observation was found related to the enhanced stability of
uranyl(V) and the absence of disproportionation in a wider pH range from acidic (pH 2) to
The effect of scan rate variations on reductive wave potential and peak current are
summarized in Fig. 5.3(ii-iv). The reductive wave potential of complexes was found to be
dependent on scan rates [Fig. 5.3(iii)] with the linear relationship between the Ecom and logν
(equation 2.30 and 2.35) and hence corroborates the redox irreversibility of UO22+-PC
complexes. The slopes of Ecom vs. logν plot for UO22+- PC complexes are given in Table 5.5.
of the cathodic shift in potential with nearly 30/αn mV per tenfold increase in scan rate [124,
180]
. Herein, the observed slopes [Table 5.5] in complexes are lower than the expected value
for Nernstian one ET reduction process. This reveals the complex kinetic and coupled
[124, 126, 180]
chemical reaction with the ET step . The decrease in peak current function
(icom/√ν) with the scan rates (CV) [Fig. 5.4(i)] and the variation in
chronopotentiometric constant (iτ1/2, where τ is the transition time for reduction wave)
with the applied currents (chronopotentiometry; CP) [Fig. 5.4(ii)] were observed in
icom/√ν and iτ1/2 quantities should be independent of scan rate (equation 2.31) and the
respectively. However, the present observation in icom/√ν and iτ1/2 quantities further
suggests the coupled chemical reaction with the ET step. Fig. 5.4(iii) describes a
typical chronoamperometry (CA) data in terms of normalized icom vs. t-1/2 related to
UO22+-PAA complex which indicates a transition between two linear regions with
151
different slopes and authenticates the association of the chemical reaction with the ET
step in UO22+-PC complexes [128, 180]. With the assumption of coupled chemical reaction,
the slope of the linear portion of Ecom vs. logν plot can be represented by equation 2.29 [124,
126, 180]
. The value of αn of complexes, calculated from their respective slope using equation
2.29 is given in Table 5.5. Fig. 5.3(ii) depict the proportional relation between the reduction
wave current and scan rates in UO22+-PC complexes. The icom vs. √ν plots for complexes [Fig.
5.3(iv)] indicates that the icom varies linearly with √ν. This is consistent with Randles-Sevcik
equation 2.31 and diffusion-limited reductive electron transfer at the electrode in UO22+-CP
complexes. The diffusion coefficient (D0) of complexes, were evaluated from their respective
slope [icom vs. √ν] using equation 2.31, and have been shown in Table 5.5. In CP, for
and kof, h obtained from slope and intercept of the linear portion of the E vs. log [l - (t
/τ) ½] curves (equation 2.33) [Fig. 5.4(iv)] are given in Table 5.5.
5.3.3.1. EXAFS study: The EXAFS experiments are conducted to get an insight into the
coordination situation around the U atom in UO22+-CP complexes. The normalized XANES
spectra of UO2CO3, U(VI)-PAA, U(VI)-PFA and U(VI)-3-PPA samples [Fig. 5.5(i)] reveal
that uranium is in the sixth oxidation state. The χ ( R ) versus R plots at U L3-edge generated
using Fourier transform from the µ (E ) versus E spectra (equation 2.17 and 2.18) are shown
in Fig. 5.5(ii) for the U(VI)-PAA, U(VI)-PFA and U(VI)-3-PPA samples. The best fitted
theoretical spectra, 1-3.2 Å in R space, along with the experimental data for the respective
samples have been shown in Fig. 5.5(ii) and the best-fit parameters have been given in Table
5.6. The correctness of the fit has been determined by the value of the R factor
defined by
equation 2.19 and mentioned in Table 5.6. The first peak at 1.3 Å and the second peak at 1.9
Å in the samples correspond to shells of axial oxygen atoms and equatorial oxygen atoms
152
respectively. EXAFS data for U(VI)-PAA was found to fit in one oxygen shell at R=2.34 Å
whereas in the case of U(VI)-3-PPA and U(VI)-PFA the data fitted in two oxygen atom shells
at the equatorial plane. In U(VI)-PAA, three oxygen atoms of coordinated water molecules
and one oxygen atom of phosphonate and carboxylate are found to be at the same distance
(R~ 2.34±0.002 Å). In U(VI)-PFA, three coordinated water molecules are obtained at a larger
distance (~ 2.41 Å) compared to chelator oxygen atoms (~2.29 Å). A similar trend was also
observed for U(VI)-3-PPA. The bond distances between U(VI) and the oxygen atom of water
molecules in U(VI)-PFA and U(VI)-3-PPA are very close to that reported in the literature
[181]
. Furthermore, the U-P and U-C distances were observed to be at 3.60±0.02 Å (N=
1.075±0.01; σ2= 0.0023) and 3.23±0.01 Å (N= 1.075±0.01; σ2= 0.001) in the U(VI)-3-PPA
respectively. These values are in agreement with the monodentate binding of phosphonate
suggests the coordination mode F (Scheme 5.2) in the present UO22+-PC complexes.
corroborate the experimental results and to find out the energetically stable molecular
structure of the uranyl(VI)-CP complexes and the reason behind their higher logβ values. The
structure optimization of complexes was tried with all possibilities of coordination modes and
it was found that the structure with chelates formation by binding through one oxygen of
phosphonate and another oxygen of carboxylate along with three water molecules in the inner
sphere of UO2 moiety are energetically more stable [Fig. 5.6(a)]. The binding energetics for
all the geometries are shown in Table 5.7. From the table, it can be observed that the free
energies of formation in aqueous complexes are very close to each other with a variation of
±10-15 kJ/mol. A similar observation was also found in experimentally determined free
energies of formation, which also has small differences with the same order of the
complexation strength.
153
The bond distances and the partial charges on individual atoms in the stable optimized
geometries and their comparison with the values in bare PCs are also estimated to understand
the complexation process at the molecular level. The bond distances between U and all the
coordinating atoms in the optimized structure of the UO22+-PCs are shown in Table 5.8. The
calculated bond distances (viz., U–Oax, U–Oeq and U–Owater) of the complexes through DFT
are nearly consistent with that by EXAFS measurements. Most surprisingly, DFT
calculations reveal an important interaction between the axial oxygen of UO22+ and the
monoprotonated phosphonate group [Fig. 5.6(a) and Table 5.8]. The intrinsic structural
UO22+ and forming stereognostic intramolecular hydrogen bond (IMHB) with axial oxygen of
UO22+ [30-31]
. This stereognostic IMHB plays a vital role in stabilizing the complexes as it
could provide an additional driving force for the coordination of UO22+ and therefore enhance
the chelating ability of PCs. The equatorial binding of PCs pushes the axial oxygens of UO22+
away from the metal ion and results in the elongation of axial O-U-O bonds. The elongation
of U-O bonds originated from axial oxygens and water molecules are visible indicators for
strong complexation of uranium by PCs and the same extent of elongation of these two bond
distances are good supporters for the nearly same logβ values of the complexes. Important
key observations can be made from partial charge calculations [Table S5.2]: (i) The charge
on uranium ion is much lower in complex form than in aquated form, (ii) The axial oxygens
have a higher negative charge in complex form than in aquated form, and the axial oxygen
involved in H-bond has relatively more negative charge than the other axial oxygen in
complex than in aquated form, (iii) The negative charge on all the oxygen atoms of both
phosphonate and carboxylate is lowered on complex formation than in free chelator, and (iv)
The positive charge on hydrogen atom involved in H-bond is slightly higher in complex form
154
The slight variations in the stability and the entropies of formation observed
complex formation. The distance between the oxygen of phosphonate and oxygen of
carboxylate (O-O) in bare PCs and complexes, the angle subtended at U by these two
coordinating atoms (∟O-U-O of complex) and the angle subtended at U by the two axial
oxygen atoms (∟O-U-O of UO2) are given in Table S5.3. The experimental trends showed
of UO2 moiety also deviated on complex formation in all the UO22+-PC systems. The extent
of deviation from linearity in UO22+ is nearly the same in all the three complexes is a further
indication for the nearly same stability of all the three complexes. The angle subtended by the
PCs atoms at central metal ion is higher in UO2–3-PPAH complex followed by UO2-PAAH
and UO2-PFAH. Thus, the chelate ring formation has a higher strain in the case of UO2-
PFAH complex and lowest in the case of UO2–3-PPAH complex. The order of the angle
subtended by PCs during UO22+ binding can be related to their spatial binding requirement.
The higher the angle subtended by CP, the higher is the spatial requirement. Therefore, the
complexes in the aqueous phase could be assigned as F (Scheme 5.2). The presence of IMHB
interaction enhances the binding affinity and selectivity for U and prevents its desorption in
and 2]−.
155
5.3.5. Aqueous speciation (stability and structure) of uranyl(V) with PCs and
[132, 184]
Spectroelectrochemical evidence: To the best of our knowledge, except carbonate
[180]
and citrate chelators, no study has been reported on the aqueous speciation of uranyl(V)
with the organic chelator. As discussed earlier (section 3.1.2), in the presence of PCs the
uranyl(V) does not undergo disproportionation from acidic pH 2 to near alkaline pH 8 in the
aqueous solution and hence it was anticipated that the uranyl(V) has aqueous redox stability
with the PCs in wide pH range. D. Cohen et al. [132] and Ikeda et al. [185] reported the stability
as the high concentration of carbonate (> 0.8 M) and limited pH range (11.7 to 12) and
suggested that any deviation from these experimental conditions will result into the brownish
precipitation. Faizova et al. [186] reported the stabilization of uranyl(V) in aqueous solution by
pH range 7-10 and the role of iron (Fe2+) binding in this stabilization phenomenon [187]
. The
present studies represent the aquatic redox stability of uranyl(V) by three PC chelators in a
broader pH range (2 to 8) with modest experimental conditions than the previous studies [132,
185]
. Herein, the appropriate reasons behind the aqueous redox stability of uranyl(V) and its
molecular parameters viz., stability (logβ) and structure with PCs were discussed.
The formal redox potential (Ef) of ET step (i) (Scheme 5.3), is described by
[180]
equation 3.3 . By substituting logβVI values of uranyl(VI)-PC complexes in
equation 3.3, the logβV values of uranyl(V)-CP complexes are evaluated and provided
in Table 5.9. The evaluated stability constant values of uranyl(V) complex with three
PCs were considerably higher than that reported by carbonate [logβ= 6.6±0.3][184] and
aqueous phase. The DFT optimized structure of the uranyl(V)-PAA complex is shown
156
in Fig. 5.6(b) and bond distances and partial charge on atoms (involved in hydrogen
bonding) of uranyl(V)-CP complexes are given in Table 5.9. Fig. 5.6(b) suggests the
of the three CP chelators with uranyl(V) and hence corroborates the proposed structure
F1 of uranyl(V)-CP complexes as shown in scheme 5.3. Table 5.9 indicates the shorter
bond distances and higher partial charges on atoms involved in IMHB interaction in
and Table S5.2]. This implies stronger IMHB interaction in uranyl(V)-CP complexes
complexes along the O=U=O axis is an important factor that provides steric protection
[188-189]
against the cation-cation interactions (CCIs) between two uranyl(V) moiety and
dimeric intermediate formation. Both the electronic and steric effects, viz., strong
disfavour the disproportionation of uranyl(V) in the aqueous phase even in the absence
supposed to be negligible in the presence of PCs due to steric protection [188-189] of oxo
[Fig. 5.7(a-b)] and 8 [Fig. S5.2(a-b)] to validate the aqueous redox stability of uranyl(V) with
PCs (as observed in the CV) and the existence of uranyl(V) species in wider pH (2 to 8) and
to acquire the electronic absorption spectra of uranyl(V)-CP complex in aquatic phase. With
the applied cathodic potential from -0.1 V to -2.6 V, the decrease in absorbance at 413 nm,
423 nm and 435 nm were observed at both pH 4 and 8. Since the absorbance at 413 nm, 423
157
nm and 435 nm are associated with the absorbance of uranyl(VI) in an aqueous medium
hence the decreasing trend with the applied cathodic potentials demonstrates a decrease in its
concentration at both pH 4 and 8, and thus the reduction process of uranyl(VI)-PAA complex.
The isosbestic points were seen at 445 nm, 453 nm, 477 nm, 484 nm, and 551 nm in
absorbance spectra obtained with applied cathodic potentials at pH 4 and 8. The presence of
[129, 131]
isosbestic points affirms the existence of a single redox equilibrium . Since the
reduction process of UO22+ to U(IV) involves the two ET steps along with U-O bond-
expected that the UO22+ to U(IV) reduction process would occur without isosbestic points.
Hence, the presence of isosbestic points in the present case validates the conversion of
uranyl(VI)-PAA to uranyl(V)-PAA species through the reduction process with the applied
cathodic potentials at both pH 4 and 8. In wavelength region 560 nm to 950 nm, the
appearance of new absorption peaks at 571 nm, 633 nm, 677 nm, 850 nm and 900 nm were
found with applied cathodic potentials at pH 4 and 8 [Fig. 5.7(b) and Fig. S5.2(b)]. These
new absorption peaks, which arise due to f-f transitions in the 5f1 configuration in the visible
these peaks have an increasing trend with the applied cathodic potentials. The assignment of
peaks positions was further corroborated by previous literature reports [129, 131-132, 180].
5.3.6. Aqueous speciation modelling of uranyl(VI) and uranyl(V) with PCs: The aquatic
performed with Hyperquad simulation and speciation (HySS) software by using their
PC along with uranyl(VI) hydroxyl species is shown in Fig. 5.8(a), which shows the
complexation process is pH-dependent and predicts the dominancy of the MLH type
158
pH >6.5, the uranyl(VI) hydroxyl species dominates over the complex species. The
speciation models of uranyl(V)-CP species along with that of carbonate and citrate are
shown in Fig. 5.8(b). From Fig. 5.8(b), it is evident that uranyl(V)-CP species is
The uranyl(V) has aquatic redox stability in wider pH range [~47 % of pH region, Fig.
5.8(c)] as well as higher logβ values with PCs than the other ligands/chelators used in
5.4. Conclusions
(molecular and redox) and binding modes of uranyl(VI) with three PC organic chelators as a
function of pH and reveal that these kinds of chelators have a very high complexation
tendency for uranyl(VI) and uranyl(V) in the aquatic environment. The comprehensive
assessed the redox variables, trends in complexing strengths and structural factors of
studies highlight the MLH type complexes and describe their thermodynamic stabilities and
complexation. EXAFS analysis and DFT calculations deduced the accurate structure as F
(scheme 5.2) of the uranyl(VI) complexes. Redox speciation studies by CV confirms the one-
electron reduction of uranyl(VI) to uranyl(V) species in the presence of PCs, provides the
complexes, and identifies their reduction pathways as PCET. The present studies describe the
uranyl(V) with PCs and validates its existence in the aquatic environment by
spectroelectrochemical measurements.
159
Schemes, Figures and Tables
160
0.06
20
0.05
Optical Density / a.u Increasing [PPA] ML
16
0.04
ε / (mol lit cm )
-1
0.03 12
M
-1
0.02 8
0.01 4
0.00 0
360 390 420 450 480 510 360 390 420 450 480 510
Wavelength / nm Wavelength / nm
(i) (ii)
100
90
etoM
80
0.142
70
tiv
ity
ity
la
60
nre
s
s
n
n
te
te
50
tio
In
In
a
40
rm
0.138
fo
30
%
20
10
0.134 0
Obs-Cal c i ntensi ty (unwei ghted)
5.0E-4
e
e
lu
lu
a
0 a
v
v
-5.0E-4
0 0.10 0.20 0.30 0.40
ti tre vol ume/ml
(iii)
Absorption spectra with increasing the addition of 3-PPA; (ii) The absorption spectra of
individual species: free uranyl(VI) and 1:1 UO22+-3-PPA complex. (iii) Hyperquad fitted
speciation diagram for the complexation of uranyl with phosphonoacetate (PAA) ([UO22+] =
161
Figure 5.2. Cyclic Voltammetry Plots of Uranyl(VI) and its complexes with PCs. (i)
[UO22+]= 10-5 M (ii-iv) [UO22+] with [PFA], [PAA] and [3-PPA] at 5x10-4 M
162
(i) (ii)
(iii) (iv)
Figure 5.3. (i) Eh-pH plots of [UO22+]= 10-5 M with [3-PPA]= 5x10-4 M, (ii) i(A) vs.
E(V) plot with varying scan rates 10-100 mV/sec and (iii) E vs. log ν plot of UO22+
163
(i) (ii)
(iii) (iv)
Figure 5.4. The different electrochemical plots of UO22+-PAA. (i) is the variations of peak
current functions (i/√ν)) with scan rates (CV). (ii) is the variation of
[1-(t/τ)1/2] plots (CP). i(A) and E(V) are considered as icom and
(CA). (iv) is the E(V) vs. log [1
Ecom.
164
(i)
(ii)
Figure 5.5. (i) Normalized XANES spectra of UO2CO3, U(VI)-PAA, U(VI)-PFA and U(VI)-
3-PPA at U L3-edge. (ii) Fourier transformed EXAFS spectra of U(VI)-PFA, U(VI)-PAA and
U(VI)-3-PPA at U L3-edge. The experimental spectra are represented by Scatter points and
165
[UO2 (PFAH) (H2O)3] [UO2 (PAAH) (H2O)3] [UO2 (PPAH) (H2O)3]
(a)
(b)
166
(a)
(b)
167
Spec iation plot of U rany l-3-PPA
100 MLH
M MH-3
etoM
80
M3 H-5
lativ M3 H-7
60
rma nre
tio
40
%fo
20 M4 H-7
MH-1
0
1 3 5 7 9
pH
(a)
(b) (c)
Figure 5.8. Speciation plots of uranyl(VI) and uranyl(V). (a) Uranyl(VI) with 3-PPA (H- is
the -OH group, M is uranyl(VI) and L is 3-PPA) (b) uranyl(V) with CP (PAA), citrate and
carbonate chelators (c) percentage of pH region of uranyl(v) stability with organic chelators.
168
Table 5.1: Protonation constants for three Phosphonocarboxylates. (Experimental conditions:
LH
7.37 ±0.08 -42.09 ±0.48 -2.54 ± 0.04 39.55 ± 0.52 132.6 ±1.7
LH2
3.16 ± 0.02 -18.05 ± 0.10 0.72 ±0.02 18.77 ±0.12 63.0 ± 0.4
LH3
1.99 ± 0.01 -11.34 ± 0.04 2.31 ±0.03 13.66 ±0.07 45.8 ± 0.2
LH
7.39 ±0.07 -42.18 ±0.40 -1.41 ±0.03 40.77 ±0.43 136.7 ±1.4
LH2
4.67 ±0.04 -26.69 ±0.23 -4.13 ±0.04 22.56 ±0.27 75.7 ±0.9
LH3
1.86 ±0.01 -10.63 ±0.05 5.70 ±0.07 16.33 ±0.12 54.8 ±0.4
LH
7.31 ±0.08 -41.76 ±0.47 0.79 ±0.01 42.55 ±0.48 142.7 ±1.6
LH2
4.56 ± 0.06 -26.03 ±0.36 0.70 ±0.01 26.73 ±0.37 89.6 ±1.2
LH3
2.07 ± 0.02 -11.84 ±0.11 8.63 ±0.08 20.48 ±0.19 68.7 ±0.6
169
Table 5.2. Assignment of Species from m/z values obtained in ESI-MS spectra.
170
Table 5.3. Thermodynamic parameters for the complexation of Uranyl with three PCs.
(Experimental conditions: T = 298 K, P = 0.1 M Pa and I =1.0 M NaClO4, [UO22+] = 0.003
M, [L] = 0.02 M; L = PFA/PAA/3-PPA).
UO22+-PAAH 12.92 ± 0.09 -73.24 ± 0.51 -4.97 ± 0.06 68.27± 0.51 224.0 ± 1.7
UO22+-3-PPAH 13.06 ± 0.07 -74.56 ± 0.40 -3.62 ± 0.03 70.94± 0.40 237.9 ± 1.3
Table 5.4. Summary of Cyclic Voltammetry data of UO22+-PCs and comparison of potential
of U(VI)/U(V) redox couple with various complexing chelators.
171
Table 5.5: Summary of electrochemical parameters obtained from fitting of Plots.
System CV: E vs. log ν CV: i vs. SQRT ν CV: Eh-pH plots CP: E(t) vs. log[1-(t/τ)1/2] plot
plot plot
Slope αn Slope D0 Region A/B Slope (S) αn Interce k0f, h
(10-8) (cm2/sec) Slope No. of pt (cm/s
(10-5) protons ec)
(10-8)
UO22+- 0.062 0.475 7.30 7.85 -0.091 ~ 1 (A) @ i= 6E-09 0.470 -0.272 1.76
PFAH -0.024 ~ 0 (B) S= 0.125
UO22+- 0.056 0.526 8.07 8.66 -0.093 ~ 1 (A) @ i= 4.5E- 0.500 -0.276 4.75
PAAH -0.015 ~ 0 (B) 08
S= 0.117
UO22+-3- 0.065 0.454 6.41 6.32 -0.093 ~ 1 (A) @ i= 3E-08 0.453 -0.296 5.04
PPAH -0.014 ~ 0 (B) S= 0.130
172
Table 5.6. Bond length, coordination number and disorder factor obtain by EXAFS fitting for
U(VI)-PCs at U L3-edge.
h r r r
σ2 0.001 σ2 0.0075±0.001
9
R-factor 0.0011 R-factor 0.016 R-factor 0.007
173
Table 5.7. The binding energy (B.E) and free energy of formation for all the three
Table 5.8. The calculate bond distances (in A0) between the uranium and the coordinating
1.790 (2.567)
1.794 (2.575)
1.795 (2.567)
174
Table 5.9: Stability constant, bond distances and partial charges of Uranyl(V)-CP complexes.
175
Chapter 6
necessity for the advancement of the nuclear fuel cycle. Within the freamework of “Green
and ionic liquids in the 21st century. Owing to its unique physicochemical properties, DES
(also americium)
um) matrices and “designer
“ solvent” for greening analysis of U in nuclear
process of americium. The present chapter is divided in three subparts (a) direct sequest
sequesting,
speciation of uranium oxides in new imidazolium based DESs, and (c) speciation of
176
6.1. Introduction:
source, is a viable option for global future energy needs because of its relatively low carbon
[1-2]
footprint . However, the prime concern of expanded use of nuclear energy are the
reprocessing and management of spent nuclear fuel (SNF) and high level waste (HLW) due
[33, 190]
to their long-term radiotoxicity and associated economic and environmental issues .
The irradiation of nuclear fuels in reactor, during electricity production, generates SNF which
contains large amounts of unused nuclear fuels, minor actinides and fission products.
Uranium (U) is primarily utilized as a fuel source for nuclear power generation and its
increasing demand as fuel material in nuclear technology activity evokes processing of metal
oxides present in naturally occurring ores as well as in spent nuclear fuel (SNF) for complete
utilization of its energy potential [33, 190-192]. Americium (241Am), a minor actinide and a major
contributor to the long-term radiotoxicity of HLW, significantly limits the storage capacity of
[193-194]
geologic repositories owing to heat production . In case of closed nuclear fuel cycle
scheme, the Am must be separated from lanthanides or curium before transmutation (i)
because its high-neutron cross sections would otherwise interrupt the fission efficiency of the
recycled fuel, and (ii) to facilitate radiologically safe fuel fabrication process [193-194].
In NFC, the current approaches of reprocessing fall into two major categories as
[34]
PUREX and pyrochemical processes . The complexities of both processes and resulting
technical challenges have opposed critically for their operation in reprocessing and demand
[35]
the development of alternative method . The first step in reprocessing involves the
dissolution of oxide fuels (mainly uranium). However, the dissolution of U matrices such as
oxides/salts/alloy is very difficult due to their poor solubility and high strength of the metal-
[39, 195-196]
oxide bond and it usually takes place in very concentrated nitric acid (~13M
HNO3). For quality assurance and nuclear forensics, the routine analysis of uranium containg
177
nuclear material samples, the sample dissolution procedure employs the time consuming
[37]
repetitive evaporation steps with hazardous concentrated acids (HNO3 and HF) . The
handling of these highly concentrated acids creates operational hurdles due to generation of
the large volume of harmful NO and NO2 gases and corrosive environment etc. Instead, if a
green dissolution is possible, it will be worthwhile to replace nitric acid in reprocessing of the
present and advanced nuclear fuels. In analysis of U, the conventional method routinely
hazardous acid waste and the involvement of various metallic impurities (Cr, Fe, Ti, Ag, etc.)
[52]
. Another option for analysis of U is the electrode modification which is very complex,
required. The Am separation from HLW is highly challenging due to its similar chemical and
through judicious ligand design is the mainstream approach for Am separation but this
approach requires multiple steps and can cause more down-stream complication due to ligand
degradation. Selective oxidation of Am(III) to its higher oxidation state is another approach
[193-194]
which could facilitate its separation from lanthanides/cm in nuclear waste streams .
green chemistry the deep eutectic solvents (DESs) have emerged as a neoteric designer green
non-aqueous media alternative to the conventional organic solvents and ionic liquids in the
178
[197-199]
DESs have been applied to diverse field of research such as electrochemistry ,
explored for the dissolution, analysis, extraction and recovery of actinides from their primary,
secondary sources, spent fuels or high-level nuclear wastes (HLW). Abbott et al. showed that
[200, 39]
malonic acid as HBD based DES viz., Maline is strongly acidic and its solvation
[200]
potential for metal oxides correlates well with that in aqueous HCl . Our interest lies in
NFC, it becomes appropriate to select the imidazolium cation as HBA since it is usually air
[204-205]
and moisture stable , fairly stable to radiolytic degradation and also its physical
properties such as density, surface tension, refractive index, etc. are not significantly affected
upon irradiation. Therefore, malonic acid as HBD and imidazolium cation as HBA are ideal
candidate for the preparation of DES and most suitable for the applications in NFC.
Literature survey implies the lack of direct sequestration of U from its solid matrices and the
greening analysis methodology for U. The present work aims to exploit Maline most
dissolution, redox and molecular speciation by IR, TGA and UV-Vis absorption,
developed to obtain the analytical figures of merit for routine U analysis using Maline as
electrocatalytic media. Interaction of Uranyl with Maline was probed by luminescence, UV-
Vis absorption and calorimetry titration. The thermodynamic interaction parameters were
evaluated to understand the driving force and mechanistic of dissolution of uranium matrices
in Maline [206]. Molecular dynamics and DFT simulations were carried out to authenticate the
experimental observations and to provide the optimized structure, free energy and molecular
orbital (MO) diagram of U species with Maline. The use of DES as co-solvent for Am
separation is reported [71] however, no studies have been reported on the speciation of Am in
179
DES. The aim of the present study is to describe the speciation of Am in two DES containing
mainly oxygen atoms as functional moiety viz., Choline dihydrogen Citrate-Malonic acid
SANEX (Selective ActiNide Extraction) processes [208] and for the growth of analytical tools
(a) U in Maline
6.3.1. Direct Sequestration and Characterization: Maline has unusual solvation potential
which can been explained based on the hole or liquid crystal theory, binding theory and
[39, 45, 209]
presence of proton as oxide-binder . According to the hole theory, the preorganized
supramolecular structure of Maline forms a polymer-like matrix, and the metal solute can
dissolve into the space (or holes) of this molecular network. Following the binding theory, the
of protons in Maline, acting as oxygen acceptor, supports dissolution through water forming
reaction with oxygen atoms of metal oxides. Therefore, in the present case, the choice of
Maline as solvating media for direct and quantitative sequestration of U matrices without any
additives. Eight U matrices: UO3, Uranyl nitrate (UN) UO2(NO3)2, UO2, U3O8, Rb2U(SO4)3,
U-metal, (U, Pu)O2 and U3Si2 and one PuO2 were dissolved using the Maline as solvating
media. The probable mechanism of the dissolution process of UO3 and related charge
[39-40, 46]
balances are given in Scheme S6.1 . The IR spectra (UO3 and UN), TGA data (UO3,
UN and UO2) and UV-visible spectra (UO3) were recorded and have delineated in Fig. 6.1 (a-
180
c). IR spectra [Fig. 6.1(a)] showed a peak near 920 cm-1 which is attributed to the stretching
frequency of [O=U=O]2+ moiety. TGA data curve [Fig. 6.1(b)] described two weight loss
region (R1A, R3) in the case of Maline while three weight loss region (R1B, R2, R3) in U
matrices dissolved Maline. This additional loss at R2 supports the dissolution process of U
matrices in Maline. The experimental and theoretical percentage loss in weight as a function
of temperature corresponding to blank Maline and U matrices dissolved Maline are shown in
Table S6.1 along with probable liberation of the various moiety. Region R1 and R3 were
assigned as decomposition of malonic acid (MA) and choline chloride (CC) moiety of Maline
[210]
with temperature respectively . The three losses R1, R2 and R3 in case of U matrices
dissolved in Maline can be assigned to loss due to choline, only H2O and H2O+HCl, and
malonate respectively. As malonic acid moiety is directly attached to UO22+ (scheme S6.1), it
is expected to get liberated at a higher temperature than choline moiety which is attached to
other end of malonic acid with very weak hydrogen bond interaction between chloride and
hydrogen atom. The region R2, observed in U matrices, may be due to the decomposition of
the equatorially attached moiety of U. In the case of UO3 dissolved Maline, the loss in region
R2 is low compared to UO2 and UN dissolved Maline, which can be due to loss of four H2O
(two from the equatorial position and other two from the solvent environment around UO22+).
In the case of UN and UO2 dissolved Maline, the loss in region R2 is due to the loss of
equatorially bonded two H2O and two HCl from the solvent. The UV-Visible absorption
peaks of species formed after the dissolution UO3 in Maline, blank Maline, and UO22+-
malonic acid complex in aqueous solution are given in Fig. 6.1(c), which reveals the
absorption peaks in the wavelength region 400-500 nm in UO3 dissolved Maline solution.
Comparison of absorption spectra of UO3 species in Maline with that of the uranium-malonic
acid complex at the same concentration of U in the aqueous medium indicates that the
absorption peaks in region 400-500 nm in UO3 dissolved Maline solution are due to the
181
complex species of uranyl [O=UVI=O]2+ with the malonic acid moiety of Maline. Regarding
(∆Amax) with respect to that of the uranium-malonic acid complex in aqueous medium and
εmax (at λmax 424 nm) were evaluated to be 10 nm, 0.23 and 18.9 M-1cm-1 respectively.
6.3.2. Speciation studies: (i) Redox speciation by electrochemistry and In situ UV-vis
viz. UO3, UN, UO2, Rb2U(SO4)3 (U std.), U-metal, U3Si2 and (U, Pu)O2 in Maline media in
potential sweep region from 1.5 V to -2.4 V were provided in Fig. 6.2(a-c). The CV of
Maline media specifies the absence of redox reaction to interfere with the U redox process in
Maline. CV of U matrices viz. UO3, UN, U std., U-metal and U3Si2 demonstrate two
successive reductive waves (peak 1 & 2) in Maline. The presence of two successive reductive
corresponding more sensitive DPV plots [Fig. S6.1(a-c)]. Redox characteristics related to
reductive waves of U matrices in Maline were proclaimed in Table 6.1. In the case of UO2
dissolved Maline solution, the prominent reductive wave was observed at nearly -1.424 V in
the CV. However, its corresponding DPV specifies two reductive waves in which the first
reductive wave is small in peak current intensity than that of UO3 and lies in the same
potential region where the first reductive wave of UO3 exit. This implies that the nature of the
first reductive wave in DPV of UO2 might be the same as that of UO3 but the concentration of
oxidative species for the first reductive wave in the case of UO2 dissolved Maline solution is
small enough compared to that of UO3. To get the clarity in nature of reductive waves of UO2
in Maline, the CV and DPV on U3O8 dissolved Maline solution were performed [Fig. 6.2(a)
and Fig. S6.1(a)]. Being mix-valent oxide, the DPV of U3O8 designates three one-electron
transfer reductive waves which could be assigned as U(VI)-U(V), U(V)-U(IV) and U(IV)-
182
U(III). The first reductive wave in U3O8 attained with nearly the same peak potential and
current intensity to that of UO3 while the third reductive wave in U3O8 [DPV: Fig. S6.1(a)]
attained at potential (-1.423 V) where the dissolved UO2 species in Maline have its prominent
reductive wave (at -1.430 V). These findings reveal that the nature of oxidative species under
reduction in the first reductive wave in UO3, U3O8 and UO2 are nearly the same and further
suggests that the nature of oxidative species under reduction at -1.430 V in UO2 Maline
would be analogous to oxidative species under the third reductive wave of U3O8 at -1.423 V.
The UVIO22+ kind of species formed through oxidative dissolution of UO3 (Scheme S6.1) [40]
in Maline (described later), therefore, it is the main oxidative species to be reduced in the first
reductive event. However, due to the slow kinetics of oxidative dissolution of UO2 in Maline,
the concentration of UVIO22+ oxidative species is small which in turn reflects in small peak
current in the first reductive event in DPV and almost negligible in the CV of UO2 in Maline
[Fig. S6.1(a) DPV; Fig. 6.2(a) CV]. Though the kinetic of oxidative dissolution process in the
case of UO2 in Maline is slow, however, it has sufficient solubility through carbonyl
UO2 species attained at -1.424 V is associated with the conversion of dissolved U(IV) species
in Maline to U(III) species. Fig. 6.2(c) and Fig. S6.1(c) exhibits the CV and DPV of PuO2
and (U, Pu)O2 in Maline respectively. As evident, the first reductive wave (peak 1) is
common in both PuO2 and (U, Pu)O2 [inset Fig. 6.2(c)] and is attributed due to the reduction
of dissolved Pu species in Maline. The second reductive wave (peak 2) obtained at -1.565 V
only in (U, Pu)O2 in CV, is further corroborated by the presence of reductive wave in
dissolved (U, Pu)O2 relative to PuO2 in Maline in DPV [Fig. S6.1(c)], is belonging to the
reduction of dissolved UO2 species of (U, Pu)O2. The dissolution of UO2 species of (U,
Pu)O2 was assumed to follow the same pattern as that was in the case of UO2 in Maline
(described above). The reductive waves (1 and 2) potential of UN, U Std., U-metal and U3Si2
183
in Maline [Fig. 6.2(b)] in their respective CV were found nearly in the same potential regions
where the UO3 keeps its peculiar reductive waves (1 and 2) respectively. This recommends
that the nature of species formed through the dissolution of UN, U Std., U-metal and U3Si2
matrices in Maline is analogous to that of UO3 in Maline. The effect of scan rates on their
peculiar reductive waves potential (Ep) and current (ip) were studied by CV and is
summarized in Fig. 6.2(d). As evident, in each U matrix, the cathodic shift in Ep and
enhancement in ip of their peculiar reductive waves with scan rates increment were observed.
The linear relationship was achieved in their corresponding ln|ip| vs. Ep-Ef and ip vs. √ν plots,
which are consistent with equation 2.34 and Randles-Sevcik equation 2.31 respectively,
matrix. For each U matrix, the values of αn and ko corresponding to their peculiar reductive
wave in Maline were evaluated from their respective slope and intercept of ln|ip| vs. Ep-Ef
plots respectively and are described in Table 6.1 along with the value of diffusion coefficient
(D0) calculated from their respective slope of ip vs √ν plots obtained by using the equation
2.31. Based on the values of αn, the two reductive waves in UO3, UN, U std., U-metal and
U3Si2 in terms of electron transfer process are described as U(VI)-U(V) conversion (Peak 1)
and U(V)-U(IV) conversion (Peak 2) and hence corroborating that these U matrices are
exhibiting two consecutive single electron transfer (SET) reduction mechanism (Scheme 6.1).
In the case of UO2 and (U, Pu)O2 the reductive wave is U(IV)-U(III) conversion. To validate
the electrochemical measurements by CV (i.e. SET reduction mechanism) and to acquire the
absorption spectral features of reduced species of UO3 in Maline, the insitu UV-vis
potentials. Fig. 6.3(i) implies the continuous decrease in absorbance at 412, 424 and 438 nm,
which are concomitant with the absorbance bands of U(VI) species (formed through
dissolution of UO3) in Maline, with decreasing potentials from no applied potential (NO V)
184
to -2.25 V due to diminution in its concentration through reduction process in its lower
oxidation states. On the other side, the growth of new absorbance peaks was observed at 512,
541, 572 and 843 nm through reduction steps [Fig. 6.3(ii-iii)]. Importantly, the incessant
increasing intensity of absorbance peaks was found at 512, 541, 572 nm but the absorbance
peak intensity at 843 nm increased initially and then decreased with decreasing potentials
[Fig. 6.3(iii)]. This kind of absorbance intensity variation at 843 nm with decreasing
potentials is reminiscent of U(V) species formation in Maline by fast kinetic of first reductive
wave upto potential -0.678 V afterwards the kinetic of second reductive wave became faster
compare to first, resulting in lowering in the concentration of U(V) species and thus
absorbance intensity at 843 nm. The incessant increase in absorbance peaks at 512, 541, and
572 nm are assigned for U(IV) species formation in Maline from both reduction steps (1 and
2) with decreasing potentials. The observed position of absorbance peaks related to U(V) and
U(IV) species in Maline are in agreement with the literature reports [129, 131, 180]. Interestingly,
[116-118]
U(V) species is quite stable and its disproportionation reaction was not observed in
Maline.
The electrochemistry of U in aqueous solvents has been widely studied earlier [98, 116-
118.148, 179-180, 211-212]
and it has been reported that the redox potential of U(VI)/U(V) couple
[98, 179, 211]
was obtained approximately at -0.250 V on GC and Pt electrode while nearly at -
[98, 179, 211]
0.2 V on Hg as working electrode . Moreover, the reported value of the diffusion
coefficient was in the order of ~ 10-6 cm2/sec [98, 179, 211]. Compare to the aqueous solvent, the
redox potential and diffusion coefficient of U in Maline deviates significantly [Table 6.1] [197-
198, 213]
. This deviation could be described in term of solvation mechanism viz. solvent
reorganization, associated free energy changes and speciation effect. The diffusing U species
in Maline media are in concentrated ionic solution thus ion-ion interactions in this media is
quite momentous and hence cannot fit into the model of non-interacting solvents. According
185
[214]
to “hole theory”, used by Abbott et al. to elucidate the diffusion of ions in DES, ionic
transport occurs only when the size of diffusing ions matches the holes within the solvent
structure. For U species, the solvation mechanism in Maline is different from that observed in
the aqueous solvent. In aqueous solvent, Marcus theory on electron transfer processes
proposed the reorganization of solvents dipoles [213]. In Maline, its constituents comprising of
hydrogen-bonded anionic complex and a bulky choline cation. The solvent reorganization in
Maline arises by the ionic translation which is quite different from the solvent dipole
ions/molecules play an important role in the reorganization process and the same has been
evaluated by molecular dynamics simulation [described later] in the present studies. The
deviation in the redox potential of U in Maline compare to aqueous solvent is expected due to
the difference in the solvent reorganization mechanism in Maline and the aqueous solvent.
[213]
The solvation structure model in two different solvents has portrayed in Scheme S6.2 .
changes and hence the change in free energy [∆G°; ∆G° = -nF(∆E°)] associated with a redox
reaction. Therefore the change in ∆G° associated with a redox reaction is affected by the
solvation structure of U species in the solvent. The shift in redox potential depends on the
extent of the solvent reorganization. The extent of solvent reorganization depends on the
reaction in the aqueous medium often disrupts the highly ordered solvation structure of water
molecules (Scheme S6.2), therefore, leading to drastic changes in entropy through solvation
of U species in the aqueous medium. In contrast to aqueous solvent, Maline has a much less
value of free energy changes (∆G°) related to redox reaction of U species occurring in Maline
186
is significantly lower. Furthermore, the speciation effect is also an important contributing
solvent due to the presence of highly coordinating functionalities as chloride and malonic
(ii) Molecular speciation by EXAFS: In aqueous solvent the chemistry of H+ and OH-
dominate the species formed in solution conversely in the case of Maline, there is
coordination competitions between chloride anion (Cl-), HBD malonic acid (MA) and H2O
for binding with U(VI) (UO22+) during the dissolution process. The coordination competitions
and structural analysis of formed species through the dissolution of UO3 in Maline have been
Maline solution (U sample) at U L3 edge is shown in Fig. 6.4(i) along with that of UO2CO3
standards. XANES spectra indicate (VI) oxidation state of U in the sample and corroborate
the absorption spectra obtained in UV-Visible spectroscopy is due to the formation of U(VI)
(oxo) species (UO22+) in Maline through dissolution. Three spectral features characteristic of
the U bond are denoted as A, B and C in the XANES spectrum. The relative position of all
three features is related to the U-Oeq and U-Oax bond lengths and their coordination number.
The axial oxygens are responsible for peak B and equatorial oxygens are responsible for peak
C. The exact peak position of all three locations is obtained by the peak fitting method, where
an arctan function is used for absorption jump and three Gaussian functions are used for peak
A, B and C. The separation between peak A and B is 9.45 eV for the U sample slightly lower
compared to UO2CO3 (9.96 eV), which indicates longer U-Oax bond for our sample. Also, the
distance between peak A and C is 38.99 eV for the U sample and 36.18 eV for UO2CO3,
which indicates slightly longer U-Oeq bonds. The χ ( R ) versus R plots, generated using
2.16 to 2.18, are shown in Fig. 6.4(ii) at the U L3 edge. A k range of 3-11 Å-1 has been used
187
for Fourier transform. Based on the XANES input, the theoretical spectra are generated for
the U structure. The best fit results are shown in Table 6.2. The first peak at 1.4 Å is the
contribution of two oxygen atoms from the axial position at 1.80 Å, confirming uranyl kind
of species [O=U=O]2+ formation via dissolution. The second coordination peak at 2.11 Å is
fitted using the two oxygen coordination shells at 2.52 Å and 2.55 Å with four and two
oxygen atoms respectively. With the HBD malonic acid component of Maline as the
chelating ligand, the existence of two dissimilar types of oxygens viz. four and two in the
primary coordination shell of U species are associated with coordination of two malonic acid
moiety of Maline through carbonyl oxygens and two H2O molecules respectively. Despite
high Cl- concentration, no U(VI)-Cl bond at distance ~ 2.74 Å [215] was observed in the U(VI)
primary coordination shell. This specifies that U(VI) exhibits exceptional coordination with
malonic acid (HBD of Maline) and H2O after the dissolution of UO3 in Maline and the
of Cl- in the primary coordination shell of U(VI) was further verified by TGA data of UO3,
which implies the wt% loss [Table S6.1] due to the equatorial attached molecule is 8% which
is corresponding to the loss of two H2O molecules from the equatorial position of U and two
from the solvent environment and hence authenticating the equatorial attachment of two H2O
6.3.3. Molecular dynamics (MD) and Density functional theory (DFT) simulations:
Initially, the bulk phase of DES (molar mixture of choline chloride and malonic acid in ratio
1:2 i.e. 500:1000) (equation 2.37) was equilibrated in NPT ensemble followed by 50 ns
production run at 298.153 K temperature and 1 bar pressure in dry condition and with 20%
water. The calculated density was found to be 1.196 gm/cc and 1.188 gm/cc for dry and
water-saturated (20%) pure DES respectively, which are in excellent agreement with the
[216]
experimental results of 1.231 gm/cc and 1.211 gm/cc respectively . The equilibrated
188
structure of the Maline DES system containing uranyl ion in the presence of water molecules
is displayed in Fig. 6.5(a). The U-Cl bond distance and U-O(water) calculated from the peak
position of the radial distribution function (RDF) [Fig. S6.2(a-b)] are 2.85 Å and 2.55 Å
respectively. The U-O(water) distance matches well with that assessed experimentally by
EXAFS. The diffusion coefficient (D0) of U in Maline was also calculated from mean square
displacement (equation 2.38). The calculated value of diffusion coefficient, 2.8x10-8 cm2/sec,
was found to be in good agreement with that obtained experimentally by CV [Table S6.1].
Using DFT simulation the structure of U in Maline has been optimized at the M06-2X/TZVP
level of theory and is displayed in Fig. 6.5(b). After obtaining a reasonable structure, next, we
deal with the thermodynamic function to evaluate the stability of the metal ion-ligand
complex. The Gibbs free energy of the complex in the gas phase was evaluated at the M06-
2X level of theory and found to be -270.3 kcal/mol. Moreover, the binding character of the
complex of UO22+ ion with MA was studied by evaluating the energy levels of MOs at the
value of 0.017 a.u. was used) is given in Fig. 6.5(c). The calculation of HB numbers against
time for pure Maline DES and in the presence of water shows that the average number of
hydrogen bonds per malonic acid molecule was found to be 0.20 and 0.18, 0.16 and 0.15 in
pure DES and in the presence of with water, respectively. It has been seen that the number of
H-bonds is decreased due to the addition of water however with more dilution (as from 10%
to 20%) there is no significant decrease in the number of hydrogen bonds. This is attributed
due to the unique hydrogen bonds during a short trajectory in Maline and hence it has a large
number of hydrogen bonding interaction and also the hydrogen bonds are present for longer
times [219]. This governs the interaction of U with Maline in the aqueous phase.
189
6.3.4. Interaction of Uranyl with Maline in aqueous phase by Luminescence and
measured by successive addition of Maline in the pure aqueous solution of UO22+ and have
been shown in Fig. 6.6(a). As evident from the figure, the UO22+ peaks intensity showing the
quenching effect with the increasing Maline concentration. The comparison study of
quenching strength of Maline and malonic acid on the luminescence intensity of UO22+ was
also carried out at the same concentration and presented in Fig. 6.6(b). The results indicate
that Maline is a better luminescence quencher than malonic acid in the aqueous phase.
Therefore, this study affirms that the diluted Maline does not resemble as one of its
constituent malonic acid and could be recognized as an aqueous chelator (different from
understand the basic chemistry behind the dissolution and then the affinity of Uranyl (UO22+)
with the Maline. In the present case, the interaction of Uranyl with Maline was probed with
show a redshift with increased intensity [Fig. 6.6(c)] on the addition of Maline in Uranyl
solution. The absorption data were fitted by using HypSpec programming to determine the
interaction strength (as log K) and to deconvolute the spectra profile to find out the
absorption profile for Uranyl and UO22+-Maline DES system [Fig. 6.6(d)] separately. The
free Uranyl ion maxima (λmax) is at 414 nm, while on interaction with DES, it was shifted to
421 nm; while the molar absorbance (ε) is changed from 7.54 mol*lit*cm-1 in free Uranyl ion
to 20.54 mol*lit*cm-1 in UO22+-Maline DES system. The thermodynamic stability (log K) for
the interaction was found to be 1.20 ± 0.01 and is relatively higher than the complexation
constants of Uranyl ion with most of the inorganic anions such as chloride, perchlorate and
190
Uranyl salts in the Maline medium and also the significant interaction of Uranyl with Maline.
The incremental addition of Maline to aqueous Uranyl nitrate solution results in thermogram
as shown in Fig. 6.6(e), which on integration would provide the heat of interaction at each
step of addition. The cumulative heat released [Fig. 6.6(f)] in the process was analysed to
calculate the enthalpy of the interaction process. The interaction of UO22+ with Maline was
found to be endothermic (∆H = 18.3 ± 0.6 kJ/mol) and is purely an entropy (∆S = 84 ± 2
J/K/mol) driven.
6.3.5. Uranium Analysis methodology: In the present case, the analysis methodology was
for quality control of nuclear fuels, nuclear material accounting and nuclear forensic
concentration of UO3 in Maline as shown in Fig. 6.7(a). The figure shows the proportional
relationship between the response peak current (ip) and added conc. of UO3. The calibration
plot [Fig. 6.7(b)] specifies the linear relationship between response peak current (ip) and
added conc. of UO3. The sensitivity, detection limit, limit of detection (LOD) and
quantification (LOQ) of calibration plot were evaluated to be 0.834 (slope of log ip vs. log
respectively. For the present methodology, the precision, accuracy and method validation
were assessed by analysing three U samples by using the calibration plot and results were
described in Table 6.3, which were further verified by other techniques viz. EDXRF and D.C.
Arc AES. The precision and accuracy of the present methodology estimated to be 0.140%
(RSD) and 0.218% respectively [Table 6.3]. Moreover, the determined concentration of U
from the present methodology is close to that of other techniques. Repeatability and
determination of U sample and corresponding RSD of precise peak currents were found to be
191
1% and 0.92%. Fig. 6.7(c), infers no interference of U reductive wave peak with other
potentially interfering metal ions (Fe, Cr, Eu, Zn, Mg, Cd, Co, Ce and Ni). Furthermore,
higher selectivity for U over other metal ions by Maline was evident from Fig. 6.7(d). The
precision (0.140%) of the present U analysis methodology is better than that of (0.2%) the
Related to the green aspects of our developed methodology, we have evaluated the
[96]
green chemistry metrics (using equation 2.49 to 2.51) of the present developed
respective green chemistry metrics have been given in Table 6.4. We have discussed the E-
factor, atom economy, reaction mass efficiency, mass intensity, quantifiable energy and
chemical consumption, and steps involved in the present developed methodology and
[220]
previous method in table 6.4. As evident from table, within the framework of green and
[221]
sustainable chemistry the developed methodology, in the present studies, indicates a
holistic approach which: (1) is cost wise economical as well as step and reagent economical
(2) uses green and sustainable Maline natural deep eutectic solvent for sequestering and
analysis of uranium matrices (3) does not generate any hazardous waste (4) has no adverse
effect on working human and environment (5) has no energy and efficiency sacrifice.
chemistry and provides a new “greener and sustainable approach” which replaces the usage
of hazardous acids as well as additives during the dissolution of uranium matrices and further
eliminates the use of hazardous reagents [acids, oxidant and reductant (use of stoichiometric
192
[220]
respectively. The present methodology follows the principles of green chemistry and
[222]
green analytical chemistry and denotes a suitable replacement for the hazardous acidic
6.3.6. Characterizations
6.3.6.1 Physical properties of imidazolium-based DES: For application of the present new
DES for dissolution, redox chemistry and electrochemical processing of uranium oxides led
Density: The density of both DES was measured in a temperature range from 10 to 85 °C and
at atmospheric pressure [Fig. 6.8 (a)]. It was observed that the densities of DES 1 and 2
decrease linearly over the studied temperature interval (equation 6.1). The density of DES 1
was observed to be slightly lower than that of DES 2. It is expected that the substitution of
the CH2 (malonic acid) atom by a heavier atom, i.e., an O atom (in diglycolic acid), results in
Õ
è <3Óä ? = g + ~ × \(é) (6.1)
Where ρ is the density in g/cm3, T is the temperature in K, and a and b are adjustable
parameters.
Viscosity: The viscosities of the selected DES were measured in a temperature range from 35
to 60 °C and at atmospheric pressure [Fig. 6.8 (b)]. The fitting parameters A (mPa·s), B (K),
and x0 (K) are shown in Fig. 6.8(b). From this figure, it can be observed that, like most ILs
and DES, the viscosity-temperature profile can be satisfactorily fitted to the Vogel-Fulcher-
Tamman (VFT) equation 6.2. The viscosity of DES 1 (604 mPa.Sec at 298 K) is lower than
193
[216,
that reported for Choline Chloride: Malonic acid (1:1) DES (1389 mPa.Sec at 298.15 K)
223]
. This indicates that DES 1 may have some ease of handling-advantage over Choline
Melting point: The thermograms [Fig. 6.8(c)] show the melting point of C10minBr, Malonic
acid and DGA as 21oC, 133.4oC and 140.3oC as against the reported values of 16oC, 135-
137oC and 140-144oC, respectively [224]. The melting point of DES 1 and DES 2 were found
to be -27.7°C and -8.7°C respectively. The peak found near 150°C is due to loss of H2O and
HBr, which is in agreement with the TGA data given in Table 6.5.
6.3.6.2 Karl Fischer Titrations: The water content of the neat DES was evaluated to be
6.3.6.3 IR spectra: Fig. 6.9(a) represented IR spectra for UO3 (red line) and UO2 (blue line)
dissolved DES along with pure DES (black line). The intensive IR bands appearing at 920
cm-1 in both UO3 dissolved DES (red line) was assigned to the antisymmetric stretching
modes of UO22+ [inset Fig. 6.9(a)]. However, interestingly, the IR bands above were found to
be absent in the case of UO2 dissolved DES (1 and 2). This showed that through the
dissolution process, probably, only UO3 got converted into a species that gave UO22+ in
[40]
solution through the oxidative dissolution process , but UO2 did not show this kind of
conversion and probably remained as U(IV) in both DES. This might be due to the absence or
6.3.6.4 TGA-DSC data: The weight percentage loss versus temperature and their
corresponding derivative plots of both pure DES and uranium oxides [UO3 and UO2]
dissolved DES are shown in Fig. 6.9 (c-d). The wt % loss vs. temperature plots indicated a
loss in two ranges of temperatures, designated as R1 and R2. The observed and calculated wt
194
% loss in both regions R1 and R2, for both pure DES and uranium oxides dissolved DES, are
given in Table 6.5. The molecules liberated during the loss are also indicated in Table 6.5.
The calculated and the observed loss in weight were found to be agreeing with each other.
The water content of DES 1 and DES 2 was found to be 15.35% and 14.8%, respectively.
This implies the association of 8.5 molecules of water for 2 building entities (2 C10mimBr+ 2
Malonic acid) of DES 1 and 8 molecules of water for 1.5 building entities (1.5 C10mimBr+
1.5 Malonic acid) of DES 2. It is also known from the IR spectra that UO3 dissolved in DES
has a UO22+ group, whereas it exists as U4+ in UO2 dissolved DES. This highly charged
uranium is associated with 4 Br- to neutralise the charge. The percentage loss in the R1
region for UO3 dissolved DES 1 is less as fewer number of HBr is liberated as compared to
UO2 dissolved DES 1. In the case of UO2 dissolved DES 1, four molecules of HBr were lost
in addition to the loss of accompanying 8.5 H2O. In the case of DES 2, the number of
associated molecules of water per molecular unit of DES was calculated to be 11. The
number of water molecules liberated from neat DES 2, UO3 dissolved DES 2 and UO2
dissolved DES 2 was found to be 5, 4 and 11, respectively in region R1. There is no loss of
HBr in the case of DES 2 in region R1, which could be due to the much stronger binding of
Br in DES 2. In contrast, HBr was lost along with other moieties in the region R2 at high
temperature in DES 2 as well as uranium oxide dissolved DES 2. The % loss of water in DES
2 is low compared to DES 1 in region R1. This indicated the higher stabilization of both
uranium oxides dissolved in DES 2 as compared to that in DES 1. The 60mg of uranium
oxides were dissolved in 1ml of DES 1 and DES 2 separately and was used in the TG-DSC
experiment. The weight loss percentage for uranium [UO3 and UO2] dissolved in DES 2 was
calculated to be 14 and 81 % in regions R1 and R2, respectively, assuming UO2 as the final
residue left behind. These values were not agreeing with the experimental values, which
195
6.3.6.5 UV-Visible spectral data: UV-Visible absorption spectra were recorded for the
solution of dissolved UO3 in DES [Fig. 6.9(b) (and inset Figures)] to understand the
mechanism of the dissolution process. For comparison, Fig. 6.9(b) included the spectra of
aqueous uranyl ion (in 1 M perchloric acid) in the presence of a higher concentration of
malonic acid (2.72 M) and diglycolic acids. It is clear from the figure that the spectral
features of UO3 [which converted to uranyl (UO22+) after dissolution in DES 1 and 2] in both
the solvents were different from that of uranyl in malonic acid and diglycolic acid in aqueous
solution [1 M HClO4] showing that the species in two cases were not the same. Importantly,
the values of molar absorption coefficient of uranyl in DES 1 (εmax=45) and DES 2 (εmax=44)
were found to be higher than those of the uranyl-malonic (εmax=6) or digcolyic acid (εmax=7)
complexes in aqueous.
dissolution steps in the case of UO3 in DES 1 are described in Scheme 6.2, leading to an
interesting finding related to the formation of uranium oxo (O=U=O)2+ species through
[40]
oxidative dissolution procedure . It may be true to propose that the dissolution steps of
UO3 in DES 2 will similarly follow the scheme. In contrast, in the case of UO2, the oxidative
dissolution process seems to be absent or slow in either of DES 1 and 2 since the probability
The electrochemical potential window of the DES was of interest and was given in
Fig. 6.9(e) along with that of the aqueous phase. It can be seen from the figure that DES 1 has
a potential window of 2.79 V while DES 2 has 2.66 V. The observed potential windows for
DES 1 and 2 were larger than those of aqueous solvent (~ 2 V). Also, both DES have an
extended cathodic potential range as compared to that of the aqueous phase and hence these
196
DES (1 and 2) are amenable and appropriate for reductive electrochemical prospects. It can
also be seen that the neat DES 1 has two reductive peaks, of its own, at 0.481 V and -1.052 V
respectively, while the DES 2 has one reductive peak at 0.373 V. These peaks did not
disappear even on long time flushing with N2 gas implying that both DES are redox-active. It
was expected that the reductive peaks observed in both neat DES might be due to the
[225-226]
reduction of imidazolium cation moiety of C10mimBr. In previous studies , the
identification of reduction products of imidazolium cation (i.e., C10mim+) has been reported
and accordingly its reduction products are N-heterocyclic carbene and subsequent dimmers.
To simulate the reduction products of DES 1 and 2 in the present case, the cyclic
voltammetry studies were performed on their reduction with varying scan rates, which infers
two SET (single electron transfer) steps in the case of DES 1 while one SET in DES 2
(scheme S6.4). In the case of DES 2, where only one reduction peak is observed is attributed
to the formation of a single product, i.e., N-heterocyclic carbine [225-226]. The DGA in DES 2,
may stabilise this due to the presence of ‘O’ in the DGA. In the case of DES 1, two SET
(single electron transfer) processes may take place. Unlike in the DES 2 case, malonic acid in
DES 1 does not have any coordinating atom in between. This makes it proceed to the second
6.3.8.2. Voltammograms
The voltammetric studies, mainly on two uranium oxides UO3 and UO2 in DES
1 and 2 were carried out. In addition, the voltammetric studies on mix-valent (VI and
[227]
IV at the ratio of 2:1) oxide of uranium viz. U3O8 was performed to corroborate
the electrochemical outcomes of UO3 and UO2 and to get an insight into the
mechanistic aspects of the dissolution procedure of UO3 and UO2 in both DES 1 and 2.
Fig. 6.10(a-b) shows the cyclic (CV) and differential Pulse (DPV) voltammetry plots
of solutions containing uranium oxides (UVIO3 and UIVO2) dissolved in DES 1 and that
197
of pure DES (without any uranium dissolved) in the potential region from 1.2 to -2.3
V. It can be seen from Fig. 6.10(a), two reductive peaks in case of UO3 at -0.305 V
(Peak I) and -1.106 V (peaks II), one reductive peak at -0.968 V in case of UO2 are
present. It also includes the data for U3O8, which has two reductive peaks (-0.295 V
and -1.150 V respectively) in DES 1. The DPV (more sensitive than CV) of UO3 and
UO2 in DES 1 was also recorded and is given in Fig. 6.10(b) which shows the trend of
the same two and one reductive peaks, respectively as acquired in their respective CV
and thus corroborate the CV results. Similarly, the CV plots of UO3 and UO2 in DES 2
are shown in Fig. 6.10(c) which describe two reductive peaks (I at -0.355 V and II at -
1.156 V) and one reductive peak (III at -1.327 V) for UO3 and UO2, respectively. The
corresponding DPV plots of uranium oxides in DES 2 are shown in Fig. 6.10(d) which
authenticates the same reductive trends as were seen in their CV plots. The peak (Ep)
and formal potentials (Ef) of reductive peaks of UO3 and UO2 in both DES 1 and 2
Another notable feature in Fig. 6.10(a) and 10(c) is the absence of oxidative peaks
in the reverse potential scan in the CV plots, suggesting the irreversibility of dissolved
species of UO3 and UO2 in both DES 1 and 2. This irreversibility in the reductive
peaks of UO3 in both DES was expected due to sluggish electron transfer kinetics
because of the required U-O bond-breaking chemical reaction after the electron
transfer steps. This reveals the formation of uranium oxo species, i.e., UO22+ through
the oxidative dissolution procedure of UO3 in both DES 1 and 2 (Scheme 6.2). On the
other hand, in CV and DPV of UO2 in DES 1, no reductive peak was observed in the
first peak potential region of UO3. This suggests that the similar oxidative dissolution
process found in UO3 was absent in UO2 in DES 1. This result is also supported by the
IR spectra of dissolved UO2 species in DES 1 [Fig. 6.9(a)] indicating the absence of
198
IR peak at 920 cm-1 of UO22+. Meanwhile, Nockemann et al. [228-229]
reported the
the functional moiety of ionic liquid with uranium ion. Following a similar concept, it
can be proposed in our case too, that the dissolution of UO2 could be through the
chelation of functional groups viz. malonic acid and bromide ion of DES 1. Now in
the case of U3O8 [Fig. 6.10(a-b)], two reductive peaks in CV and DPV plots were
observed in the same potential regions as of UO3 which reinforces our idea that U3O8
in DES 1 too follows the oxidative dissolution process similar to UO3 and the
formation of similar kinds of species i.e., UO22+. In the case of DES 2, the DPV of
UO2 [Fig. 6.10(d)] reveals the reductive peak I (at -0.162 V) with low current intensity
in the same potential region where the UO3 has its first reductive peak (-0.169 V).
However, it is to be noted that, unlike the DPV, the CV of UO2 did not show any
reductive peak in that potential region [Fig. 6.10(c)]. The above small peak at -0.162 V
in DPV is related to the redox activity of dissolved species of UO2 in DES 2 and is
probably because of more sensitivity of DPV than CV. Further, the appearance of
small current intensity [Fig. 6.10(d)] suggests the slow kinetics of oxidative
plot [Fig. 6.10(c)], the dissolution through DGA chelation was assumed as a major
conversion process in the case of UO2 in DES 2, which was corroborated by the DPV
plot of U3O8 in DES 2. The DPV of U3O8 in DES 2 [Fig. 6.10(d)], indicates three
reductive peaks at -0.171 V (I), -0.898 V (II) and -1.232 V (III). Since U3O8 is a
[227]
stoichiometric mixture of two UO3 and one UO2 , it can be noted as follows: Out
of the three reductive peaks, the two reductive peaks of U3O8 occurred at nearly the
same respective potentials of reduction of UO3 (I and II) and the third reductive peak
of U3O8 occurred at the same potential as that of one single reductive peak of UO2
199
(III). Hence, the dissolution process of U3O8 in DES 2 could be the result of both
dissolution mechanisms as envisaged in UO3 and UO2, i.e., in addition to the oxidative
functional groups (viz. DGA and Br-) of DES 2. Therefore, the finding of reductive
peaks I, II and III in the case of U3O8 validates two different mechanisms of
reductive peak of blank DES 1 and that of UO3 (peak II) as they occur at almost
0.902 V in DPV and at -1.106 CV) [Fig. 6.10(a-b)] which created ambiguity during
the identification of these reductive peaks. This was resolved by recording the DPV
plots with increasing concentration of UO3 in DES 1. It also gave a value addition to
this exercise which established a method for the analytical determination of uranium
indicated the enhancement in response peak current (ip) of reductive peaks I and II of
UO3 with increasing UO3 concentration in DES 1 [Fig. 6.11(a-b)]. The linear
relationship (Adj. R-Square =0.997 and 0.991 and slope= 2.131x10-6 and 3.962x10-6
A/mg.ml-1 for reductive peak I and II respectively) between the response peak currents
(ip) and concentration of UO3 was observed for both reductive peaks [Fig. 6.11(c-d)].
The change in potential (Ep) of reductive peaks I and II with varying concentration of
UO3 is also examined and two linear region with different slopes was observed [Fig.
6.11(e)]. At point A, the reductive species is blank DES 1 itself, whereas the region
BC is related to the reductive peak I and II of UO3 species in DES 1. Therefore, the
findings of Fig. 6.11(a-e) clearly tell us that the reductive peaks I and II corresponding
to UO3 in DES 1 in CV and DPV [Fig. 6.10(a-b)] are indeed due to the reduction of
200
uranium species in DES 1. Instead of analyzing separately UO3 and UO2 in DES 2, it
was simplified with a similar exercise for U3O8 in DES 2 (because U3O8 gives three
peaks pertaining to both UO3 and UO2) and the results are given in Fig. 6.12(a-b).
Similar to the above case of UO3 a linear relationship was obtained [Fig. 6.12(b)]
between response peak current (ip) and concentration of U3O8 corresponding to the
reductive peaks I to III. The slopes of ip vs. concentration of U3O8 plots were found to
be 1.44x10-6, 7.32x10-7 and 9.03x10-7 A/mg.ml-1 for reductive peaks I, II, and III,
respectively.
The calibration plots [Fig. 6.11(c) and Fig. 6.12(b) (i)] related to uranium
uranium oxide sample with the present voltammetry method and with another
was evaluated from their respective calibration plots [Fig. 6.11(c) and Fig. 6.12(b) (i)]
and are given in Table 6.7. Also, the concentration of uranium oxide samples analyzed
by the biamperometry method is reported in Table 6.7 and found to agree with those
6.3.8.4. Scan rate variation and determination of transport and kinetic parameters
The scan rate variation was studied for solutions of UO3 and UO2 dissolved in both
DES to find out the diffusion coefficient (D0) and kinetic parameters (k0, αn). The CV
plots of reductive peaks of uranium oxides (UO3 and UO2) in both DES at various
scan rates are recorded and are shown in Fig. 6.13(i-ii) for DES 1. The electrochemical
equation 6.3 was applied to calculate αn at each scan rate. The value of αn is < 0.5 for
UO3 and UO2 in both DES. Fig. 6.13(i-ii) demonstrates the shift in peak potential (Ep)
201
and reductive peak current (ip) enhancement on going from lower to higher scan rates.
[121-122]
Randles-Sevcik equation (equation 2.31) was applied to estimate the diffusion
º.¶¿µ °±
|® L ®/ | = (6.3)
´²³
Where Ep is the peak potential, Ep/2 is half peak potential, F is the Faraday constant, α is the
charge transfer coefficient, n is the number of electrons transferred. The plots between ip
uranium oxides in both DES [Fig. 6.13(i-ii)]. Therefore, it can be concluded that in
both DES, the UO3 and UO2 undergo reduction through a diffusion-controlled process.
From their respective slopes of linear plots above, the diffusion coefficient values (D0)
of uranium oxides (UO3 and UO2) in both DES were calculated and reported in Table
6.6. It can be seen from the table, the order of D0 for UO3 is DES 1> DES 2 and
similar is the case with UO2. This shows that in DES 2, the formation of bulkier
uranium species is more likely. However, when we compared the oxides of uranium in
a particular DES, D0 of UO2 > D0 of UO3 in DES 1 but in DES 2 the order is reversed.
As discussed earlier, the irreversibility of the reductive peaks of UO3 and UO2
in both the DES [Fig. 6.10(a) and 10(c)] can further be validated with the linear
[75-76]
relationship between Ep and log ν according to the equation 2.35 . The Ep vs. log
ν plots for the reductive peak of UO3 and UO2 are presented in Fig. 6.13(i-ii). All these
plots show the linear relationship between Ep vs. log ν, corroborating the said
irreversibility of uranium oxides in DES 1 and 2. The value of the formal potential (Ef)
required to calculate αn and k0 was deduced from the intercept of Ep versus ν plot (not
given here) on the ordinate by extrapolating the line up to ν= 0. The values k0 can be
calculated from the above equation 2.35. However, in order to verify the values of k0,
other equations reported in the literature were also used as discussed below.
202
According to the literature, for irreversible redox process, an expression of ip in
[75-76]
terms of Ep can also be described by the equation 2.34 . A plot of lnip versus (Ep-Ef)
would give a straight line with a slope of (αnF/RT) and intercept of ln (0.227 nFAC0*k0). The
corresponding lnip vs. (Ep-Ef) plots of uranium oxides are provided in Fig. 6.13(i-ii), which
illustrate the linearity between lnip versus (Ep-Ef). The values of αn were calculated from the
slopes of the above linear plots (equation 2.34) for UO3 and UO2 in both DES and are
reported in Table 6.6. As can be seen from the table, the values of αn are < 0.5, corroborating
the sluggishness of the electron transfer process. Substituting the known values for the
respective terms in intercepts from the above plots, one can obtain the values of k0. However,
its determination using the above two equations (2.34 and 2.35) depends on the knowledge of
the value of Ef which is not so easy to obtain for irreversible redox conditions. In those cases
where the calculation of Ef is not so simple, it is derived from the literature (Velasco et al.)
[92]
that equation 2.36 can also be used, eliminating the Ef, for the calculation of k0. Knowing
the values of Ep-Ep/2 at a particular scan rate and with the knowledge of D0 (as
determined above), equation 2.36 can be used to calculate k0. The values of k0 were
evaluated using equations 2.34, 2.35 and 2.36 for UO3 and UO2 in both DES and are
mentioned in Table 6.6. From table, it is clear that the values of k0 calculated from are
found to be nearly the same except that of UO3 in DES 2. The lower values of transfer
coefficient of uranium oxides in both DES, in the present case, indicate the asymmetry
of the configuration during the electron transfer between the initial and final
configurations and also the difference in the mechanism of electron transfer in DES
The deviations were observed in redox parameters (viz. peak potential (Ep),
transfer coefficient (α) and diffusion coefficient (D0)) of dissolved uranium species in
203
both DES from that of the aqueous solvent. For example, the redox potential values of
the U(VI/V) redox couple lies in the range -0.2 V to -0.250 V at Hg, glassy carbon
(GC) and Pt working electrode respectively [98, 116-118, 148, 179-180, 211-212]. However, in the
present studies, the same redox couple's peak potential values were observed at -0.305
V and -0.355 V in DES 1 and 2 respectively at the GC electrode. Also, the values of
the diffusion coefficient of uranium species [10-8 cm2/sec] in DES were seen to be
lower than that in aqueous solvent [10-5-10-6 cm2/sec]. This deviation could be
structure of the uranium ion is highly ordered in the aqueous solution while the same
[213]
is expected to be less ordered in the case of present DES . Therefore, the free
energy change (∆G°= -nF∆E°) during a redox reaction of uranium species in the
Accordingly, the difference in redox potential values of uranium species was observed
in two different solvents (viz. aqueous and DES). Also, the speciation of uranium in
of UO3 and UO2 in DES 1 and 2. Further, it was aforementioned that UO3 converted to
UO22+ [U(VI)] through dissolution in both DES. Therefore, the plausible redox
mechanism for UO3 in DES 1 and 2 can be described by scheme 6.3. Two consecutive
SET (Single electron transfer) reduction mechanism was anticipated in case of UO3
dissolved in DES 1 and 2 as discussed in scheme 6.3. In the case of UO2 in both DES,
the reduction mechanism is proposed as one SET through U(IV) to U(III) redox
conversion [231].
204
6.3.9. In situ spectroelectrochemical Studies
UO3 dissolved DES 1 and 2 solutions for preliminary understanding of the redox stability of
the uranium oxidation states and acquiring the electronic absorption spectra of intermediate
redox species of uranium during electrolysis. Fig. 6.14(a-c) depicts the resultant absorption
spectra recorded at different applied potentials of UO3 in DES 1. As clear from Fig. 6.14(a-c),
in applied potential range from 0.45 to -1 V, the observed spectra no longer passed through
the isosbestic point, indicating the absence of the single redox equilibrium. This observation
validates our earlier CV result [Fig. 6.10(a)], where two redox processes were observed. The
decrease in absorbance values of UO22+ [U(VI)] at 422, 428, 455, 463 and 480 nm was found
with the application of cathodic potential from 0.45 V to -1 V [Fig. 6.14(a) and its inset
figure]. This indicates the conversion of U(VI) to its lower oxidation states. The decrease in
the concentration of U(VI) with cathodic potential was reflected in terms of its absorbance in
the spectra. Fig. 6.14(b) shows the enhancement in absorption intensity and change in the
nature of spectra (regions AB and BC) in the range of the applied cathodic potentials from
than -0.042 V, a decreasing trend in absorbance intensity was found. The increasing and
decreasing absorption intensity trends are associated with the increasing and decreasing
formation of intermediate redox species. Since the one absorption peak at 780 nm was
observed in Fig. 6.14(b), it was assigned to intermediate species, i.e., U(V) in DES 1. The
assigned peak position of absorption spectra corresponding to the intermediate U(V) species
(transient) obtained during electrolysis is in agreement with the previous literature reports [129-
131, 231]
. The U(V) formed this way gets further reduced to U(IV). Therefore, the concentration
of U(V) increases initially because of the fast kinetics of the conversion [U(VI)-U(V)] due to
205
the first reductive peak up to -0.042 V. Beyond this, the kinetics of the conversion of the
second reductive peak [U(V)-U(IV)] becomes faster than the first, resulting in the decrease in
U(V) concentration. Also, it is clear that with applied cathodic potentials, the concentration of
U(IV) species continuously increases. Hence the absorption peaks at wavelengths 569 nm,
609 nm, 630 nm and 646 nm in Fig. 6.14(c), which shows the enhancement in absorption
intensity with applied cathodic potentials from -0.042 V to -1 V, are due to the U(IV) species
obvious from Fig. 6.15(a), with applied cathodic potentials from 0.0419 V to -2.79 V, the
spectral changes were observed with an isosbestic point at 482 nm, indicating the existence of
[129-131]
one redox equilibrium during the region of applied cathodic potentials . Absorbance
intensities at 421, 428, 445, 457 and 466 nm were continuously decreasing with the applied
cathodic potentials from 0.0419 V to -2.79 V. Obviously, these absorption peaks are
associated with the UO22+ [U(VI)] species in DES 2. Hence the decreasing order is due to the
lowering in the concentration of U(VI) species with applied cathodic potentials. In the NIR
region (>725 nm), the absorbance intensities in regions AB and BC at 774, 795, 841, 850,
865, 904 and 921 nm [Fig. 6.15(b)] were found to be dependent on applied cathodic
potentials and show the increasing trend. Literature reports [129-131, 180, 231] suggest the presence
of U(V) species in the NIR region at the above wavelengths. Therefore, in Fig. 6.15(b), the
absorption peaks observed in the above wavelength regions (i.e., AB and BC) are assigned to
U(V) species in DES 2. The species of U(V) is relatively stable (as lower than -1.4 V, no
peaks corresponding to U(IV) is present) in DES 2 and increasing in concentration with the
varying applied cathodic potentials. Therefore, the absorption peaks of U(V) species are
highly intense even when the cathodic potential was varied up to -2.79 V. In Fig. 6.14(b), in
DES 1, only one absorption peak of U(V) was obtained at 780 nm due to the formation of
206
intermediate species (transient) of U(V), while several absorption peaks (at 774, 795, 841,
850, 865, 904 and 921 nm) were observed in DES 2 [Fig. 6.15(b)] due to the formation of
Molecular dynamics (MD) and DFT: The MD simulation was performed to corroborate the
in DES and obtaining the probable equatorial coordination of dissolved uranium species
through radial distribution function. Taking the simulation of UO22+ ion, as an example,
which formed after dissolution in DES 1 [discussed above], first the geometries of C8mimBr
(C8 instead of C10 is taken here, as the basis set was already available for C8 and assumed to
apply for C10 also) and malonic acid (MA) were first optimized using M06-2X functional [84]
[85] [86]
and TZVP basis set as implemented in Turbomole suite of package . The atoms of
C8mimBr and malonic acid are represented in Fig. 6.16(a). Initially, the bulk phase of DES
(molar mixture of C8mimBr and malonic acid in ratio 1:2 i.e. 500:1000) was equilibrated in
NPT ensemble followed by 50 ns production run at 298.153 K temperature and 1 bar pressure
in dry condition and with 20% water. The calculated density of C8mimBr was found to be
[232]
1.14 gm/cc which is in excellent agreement with the experimental results of 1.16 gm/cc .
The system containing UO22+ ion was equilibrated in NPT ensemble followed by 50 ns
production run at 298.153 K temperature and 1 bar pressure. In order to study the
coordination environment of the UO22+ ion in DES solution, the enlarged structure of the
complex of UO22+ ion is presented in Fig. 6.16(b). The U-Br bond distance is calculated from
the peak position of the radial distribution function (RDF) of U and Br as displayed in Fig.
S6.3(a) and the values are presented in Table 6.8. The U-Br bond distance was 2.85 Å. The
U-O (O of malonic) bond distance was calculated from the peak position of the radial
distribution function (RDF) of U and O of malonic acid as displayed in Fig. S6.3(b) and the
207
values were presented in Table 6.8. The malonic acid was shown to be monodentate, and only
carbonyl O participated in the first coordination sphere towards the UO22+ ion. The calculated
U-O bond length was found to be 2.55Å. The average number of malonic acid molecules in
the first coordination shell was found to be ~2 [Fig. 6.16(b)]. The diffusion coefficient of
UO22+ ion was calculated from mean square displacement and was found to be 1×10-8
agreement with the experimentally evaluated value (1.016×10-8 cm2/sec) (UO3, peak I; Table
6.6). Further, in order to understand the interaction of malonic acid with uranyl ion in the
presence of Br- ion, the structure of uranyl ion with malonic acid has been optimized at the
M06-2X/TZVP level of theory and is displayed in Fig. 6.16(c). The Gibbs free energy of the
complex in the gas phase was evaluated at the M06-2X level of theory and the same is listed
in Table 6.8. Moreover, the binding character of the complex of UO22+ ion with MA was
studied by evaluating the energy levels of MOs at the B3LYP/TZ2P/ZORA level of theory
[217-218]
. The Molecular orbital (MO) diagram of LUMO, LUMO+1, HOMO and HOMO-1 of
the metal-ligand bonding for [UO2-(MA)2-(Br)2] complex (isosurface value of 0.017 a.u. was
not thermodynamically simple and is hampered by high potential (2.6 V vs. SCE) for Am
(III)/Am(IV) couple according to Latimer diagram, which is shown in Fig. 6.17(i)(a). In the
present study, the speciation of americium oxide in CDHC-MA DES is performed using
oxidation state, especially in its actinyl (AmO22+ or AmO2+) ion. Following a direct approach,
the 5 mg of Am-oxide was successfully dissolved in CDHC-MA DES without any additives
208
using dissolution set-up in glove box (discussed in chapter 2). The change in colour of
solvent (yellow colour in Fig. 6.17(i)(b)) and γ-counting through NaI(Tl) and HPGe detector
authenticates the solubility of Am-oxide and its presence in CDHC-MA DES through
dissolution process. The UV-Visible absorption spectra of dissolved Am-oxide sample was
recorded, which shows the dominant absorption peak at 502 nm. This indicates the formation
carried out by cyclic (CV) and differential pulse voltammetry (DPV). The representative CV
plots in potential region form -0.2 V to 1.5 V at Pt working electrode are shown in Fig.
6.17(ii)(a). The initial results indicate that (i) no interference from blank DES on the
oxidation process of Am(III), (ii) two oxidation peaks with one electron transfer as Am(III) to
Am(IV) and Am(IV) to Am(V) was observed and hence Am(III) can be easily oxidized to its
higher stable oxidation state, (iii) Am(IV) is a stable oxidation state in CDHC-MA and
Am(III)-Am(IV) conversion has much lower oxidation potential (~ 0.68 V) in DES compared
to that in aqueous medium (i.e., 2.6 V vs. SCE). The DPV plot [Fig. 6.17(ii) (b)] with
increasing concentration of Am(III) implies that the oxidation peak II i.e., peak belonging to
AmO2+ formation is relatively unstable. This may be due to the combined effect of three
depends on AmO2+ concentration and (iii) requirement of Am-O bond formation during
Am(III) to AmO2+ oxidation steps, which is kinetically sluggish. In CV, the scan rates (ν)
variations [Fig. 6.17(ii)(c)] indicates enhancement in peak current of both oxidation peaks
with increasing ν and the linearity between peak current (ip) and square root of scan rate (√ν).
This implies the diffusion-controlled oxidation process of Am(III) in CDHC-MA. From the
slope of ip vs. √ν plot and using Randles-Sevcik equation (equation 2.31), the diffusion
209
coefficient of Am in CDHC-MA was calculated and reported in Table 6.9. The
was evaluated using equation 2.34 and reported in Table 6.9. Theoretical calculations
were performed to corroborate the electrochemical results and to obtain the optimized
structure and energetic of Am in both DES. In order to understand the interaction of CDHC-
MA with Am, structure of their complexes have been optimized at the B3lYP/SVP level of
theory and is displayed in Fig. 6.17(i)(d). Further, free energy of complexation in solution
phase was determined using the total energy of the chemical species at the B3LYP/TZVP
level of theory using standard thermodynamic corrections and COSMO solvation model [233].
The calculations reveal that 1:2 type of Am-DES complex formation and their corresponding
6.4. Conclusions:
This study affirms that Maline is a versatile all in one platform working as dissolver
medium, designer solvent for analysis of U in nuclear materials and promising U chelator.
Present studies reveal that Maline and Imidazolium based DESs have unique solvation
potential for diverse kinds of U matrices like oxides, salts, metal powder and alloy without
any additives and hence these DESs could be acknowledged as low temperature direct
sequestering media for processing of U matrices to incorporate U ion in solution. The density,
viscosity and melting point of newly prepared imidazolium-based DES 1 and 2 are reported.
The neat DESs and dissolved uranium matrices have been characterized by IR, UV-Vis,
TGA, Karl-Fischer titrations etc. The present studies indicate that DES 1 and 2 are redox-
active in nature. The redox speciation of U matrices in Maline and imidazolium based DES 1
and DES 2 were successfully probed by electroanalytical techniques viz. CV and DPV. The
redox thermodynamic characteristics (peak potentials (Ep) and formal potential Ef), diffusion
210
coefficient (D0) and kinetic parameters (k0 and αn) of uranium species in above DESs were
determined and compared with the literature. The EXAFS analysis specifies the UO22+ kind
of species formation with malonic acid and H2O at equatorial coordination through
and provide the electronic absorption spectra of U(V) and U(IV) species in Maline and
imidazolium based DESs. MD and DFT simulations showed very good agreement with the
experimental interpretations and deliver the optimized geometry, binding energy and MO
diagram of U species in above DESs. The interaction of the UO22+ solution with Maline was
observed to be endothermic and entropy-driven and logK value was found to be 1.20 ± 0.01.
The redox thermodynamic and kinetic parameters of Am redox process and structure of Am
211
Schemes, Figures and Tables
Part (a) U-matrices in Maline
212
(a)
(b) (c)
Figure 6.1. Characterization: (a) IR spectra of UO3 and UN dissolved Maline solution. TGA
plots (b) Wt% loss versus Temperature (c) The UV
UV-Visible
Visible absorption spectra of UO3
dissolved in Maline solution, blank Maline and UO22+ and malonic acid complex in aqueous
solution.
213
Figure 6.2. Redox speciation of Uranium matrices in Maline. Cyclic voltammetry plots of (a)
UO3, UN, UO2, U3O8 and blank Maline (b) U std., U metal and U3Si2 (c) PuO2 and (U, Pu)O2
(d) Scan rates variation studies by Cyclic voltammetry (CV) on UO3 dissolved in Maline.
214
(i)
(ii)
(iii)
Figure 6.3. Insitu UV-visible Spectroelectrochemical Data. (i) Absorbance spectra of U(VI)
in Maline with applied cathodic potentials. (ii) and (iii) are the appearance of new absorption
peaks with applied cathodic potentials related to U(IV) and U(V) species in Maline
respectively.
215
(i)
(ii)
Figure 6.4. (i) Normalised XANES spectra at U L3-edge along with standard samples. (ii)
Fourier transformed EXAFS spectra of U sample at U L3-edge in (a) R space and (b) K space
along with the best fit. The experimental spectra are represented by Scatter points and the
216
(a) (b)
-10.0 120A
119A
118A
-10.5
-11.0
Energy level (eV)
-12.0
-13.0
-14.0
-15.0 117A
-15.5 116A
-16.0 115A
[UO2-(MA)2-(H2O)2]2+ complex
(c)
Figure 6.5. (a) Equilibrated structure of the Maline system containing Uranyl ion in the
presence of water molecule. System representation with 5% water. color code: Steel: UO2
species; Red & White: H2O; Cyan: Cl; rest are malonic acid and choline (b) Optimized
structures of [UO2-(MA)2-(H2O)2]2+ (c) The calculated energy levels (eV) of the MOs of the
217
Figure 6.6. (a) The luminescence intensity of Uranyl with increasing concentration of Maline
(b) Comparison of luminescence intensity of Uranyl with Malonic acid and Maline at two
different concentrations (c) The absorption spectra for Uranyl titration with Maline (d)
Deconvoluted spectra representing the absorption profile for the pure components (e) Raw
calorimetric data for the interaction of Uranyl with Maline (f) Integrated Data showing the
cumulative heat released on successive additions of DES solution to fixed volume of Uranyl
solution; [UO22+] = 0.01 M, [DES] = 0.48 M.
218
Figure 6.7. (a) Differential pulse voltammetry plots with increasing concentration of UO3 (b)
Calibration Plot of UO3 dissolved in Maline solution obtained with increasing concentration
(c) Electrochemical interference study on UO3 dissolved Maline with other metal ions
dissolved Maline (d) Selectivity of Maline for different dissolved metal ions.
219
Table 6.1: Electrochemical thermodynamic and kinetic parameters of U matrices in Maline.
220
Table 6.2: Bond length (R (Å), phase-corrected), coordination number (N) and disorder
Table 6.3: Analytical figures of merit and validation of developed analytical methodology for
(mg/ml)
(Unknown conc.)
(Unknown conc.)
221
Table 6.4: Comparison of Green metrics of present methodology (direct sequestering and
analysis of uranium matrices with Maline supramolecular scaffold) and previous method (i.e.
hazardous acidic/additives dissolution and biamperometery analysis).
(i) and (ii) are the conditions without solvents and with solvents respectively. More details calculations are
given in supporting information.
1a and 1b is the atom economy in the case of UO2
2 is the atom economy in the case of UO3
222
Part (b) U-oxides
U in Imidazolium based DESs
223
Figure 6.8. (a) measured density of DES 1 with varying temperature and fitting (b) measured
viscosity of DES 1 and fitting (c) Thermograms obtained in DSC for pure components.
224
Figure 6.9. (a) IR spectra of the UO3 and UO2 dissolved in DES 1 (b) UV-Vis spectra of UO3
dissolved DES 1 (c) TGA data of the UO3 and UO2 dissolved in DES in terms of the wt %
loss with temperature (d) derivative plot (e) electrochemical potential window and redox
peaks of DESs.
225
(a) (b)
(c) (d)
Figure 6.10. (a) and (b) are the CV and DPV plots for the oxides of uranium dissolved in
DES 1. (c) and (d) are the CV and DPV plots for the oxides of uranium dissolved in DES 2.
226
Figure 6.11. (a) and (b) are the DPV plots with increasing concentration of UO3 (peak I and
II) in DES 1. (c) and (d) are the ip(A) versus conc. UO3 plots in DES 1 (e) Ep vs.
227
(a)
(b)
Figure 6.12. (a) is the recorded DPV plots for U3O8 by varying its concentration (DES 2). (b)
(i) to (iii) are the ip(A) vs conc. U3O8 plots for reductive peaks I to III.
228
(i)
(ii)
Figure 6.13. (i) and (ii) are the scan rates variation studies on uranium oxides dissolved in
DES 1 in cyclic voltammetry (CV). (a) is the CV plots recorded with varying scan rates. (b)
is the E vs. logν plots. (c) is the ln|ip| vs. Ep-Ef plots (d) is the i(A) vs. √νν plots of their
229
(a) (b)
(c)
U(VI) in DES 1 with applied cathodic potentials. (b) and (c) are appearances of new
absorption peaks with applied cathodic potentials related to U(V) and U(IV) species in DES 1
respectively.
230
(a)
(b)
U(VI) in DES 2 with applied cathodic potentials with the isosbestic point. (b) (i) to (iii)
showing the appearance of new absorption peaks with applied cathodic potentials related to
231
Figure 6.16. MD simulation results: (a) Geometries of (i) C8mimBr and (ii) malonic acid
molecules. (b) Enlarged picture of DES-uranyl system. (i) UO2-Br2, UO2-Br3 and UO2-Br4
and (ii) UO2-2malonic acid. Here, U: pink; O: Red; Br: light blue; malonic acid: Blue. DFT
simulation results: (c) Optimized structures of [UO2-(MA)2-(Br)2] (d) The calculated energy
levels (eV) of the MOs of the [UO2-(MA)2-(Br)2] complex using B3LYP/TZ2P/ZORA level
of theory.
232
(i)
(ii)
Figure 6.17. (i) (a) Latimer diagram for Am in 1 M perchloric acid (b) dissolved sample of
Am-oxide in DES (c) UV-Visible absorption spectra of Am in DES (ii) (a) CV plots of Am-
oxide in CDHC-MA and blank CDHC-MA (b) DPV plots with increasing concentration of
Am in CDHC-MA (c) CV plots with varying scan rates from 10 mV/sec to 150 mV/sec (d)
233
Table 6.5: The observed and calculated weight loss along with the liberated molecules at various temperatures in TG-DSC.
Solution
of UO3 in
DES 1 29.5 25.4 23-179 2HBr+8.5H2O 68.2 52.75 179-414 2C10mim+2Malonate
Solution
of UO2 in All the remaining
DES 2 elements after liberation
18.2 11.0 25-188 11.0 H2O 79.6 73.6 188-358
of 11 H2O leaving UO2
as final product
234
Table 6.6. Redox thermodynamic, Diffusion coefficient and kinetic parameters of UO3 and
UO2 in DES (# eqn-5; *eqn-6; @
eqn-7. PI is the reductive peak I. PII is the reductive peak II).
parameters
DES-1 DES-2 DES-1 DES-2
Peak Potential -0.305 (CV) (PI) -0.355 (CV) -0.968 (CV) -1.327 (CV)
(Ep) (V) -0.128 (DPV) (PI) -0.167 (DPV) -0.866 (DPV) -1.230 (DPV)
(mg/ml) (mg/ml)
235
Table 6.8. The calculated values of structural parameters in simulated systems and the
calculated values of Gibbs free energy (kcal/mol) of complexation in the gas phase.
236
Chapter 7
Aim: This chapter summarizes the thesis, discusses its findings and contributions, points out
limitations of the current work, and outlines directions for future research.
ions especially U, Np and Am were carried out with aqueous soluble organic chelating
ligands and deep eutectic solvents media using electrochemistry, UV-Visible absorption
The summary of works presented in the thesis is divided into two parts:
(a) Speciation of actinyl ions in aqueous soluble chelating ligands: The studies
transformations and kinetics information about the actinyl species (U, Np and Am) in the
aqueous. The speciation of uranyl and neptunyl was discussed with chelating ligands viz.,
237
decontamination process or potentially present in waste repository scenarios. The studies in
the present thesis infer that citrate, PPA and PCs chelators form very stable aqueous soluble
complexes of U and Np, which are quite persistent over a wide range of Eh-pH. The stability
constant (logβ), redox potential and shift in redox potential of uranyl with these chelators
were found to be as citrate<PPA<PCs (Fig. 7.1 and 7.2). The hetero-bidentate chelation in
complexation tendency of PCs towards UO22+. Redox speciation of actinyl ions is probed to
deduce redox energetic and kinetics of electron transfer process for their redox couples in
aqueous solution across the pH range. In aqueous solution, the peak potentials of actinyl ions
(U and Np) were measured and the potentials were compared in absence and presence of
chelators. To obtain their similarities and differences in terms of the shape of the cyclic
PPA and PC ligands have major impact on the U and Np redox system in terms of reaction
couple chemical reaction was discovered with electron transfer (ET) steps for UO22+
complexes, which suggesting reduction mechanism as ICI and PCET in citrate and PCs,
respectively. Interestingly, the rare oxidation state of uranium i.e., pentavalent uranyl (UO2+)
species was observed to be redox stable in a wider pH in presence of citrate and PCs in
aqueous solution and its validation for existence, electronic absorption spectra, stability
and DFT. This information might be important in detailed understanding of actinide (V)
chemistry.
238
Figure 7.1. Variation of stability constant (logβ) of actinyl ions with various complexing
ligands
Figure 7.2. Variation of shift in peak potential (|∆E|) (logβ) of uranyl ions with various
complexing ligands.
(b) Dissolution and speciation of uranium matrices and americium oxide in actinide
specific deep eutectic solvents: Deep eutectic solvents namely Maline and imidazolium
cation based DESs shows better solvation potential for diverse kinds of actinide matrices. For
the first time, two-imidazolium based DESs were prepared by combining the ionic liquid and
239
carboxylic acid and being reported on the dissolution of uranium oxides. Above DESs
solvents favour the dissolution process of solid uranium samples through their oxygen-
binding sites. Malonic acid based DESs viz., Maline and CDHC-MA works as designer
solvents for (1) quantitative green analysis of NMs such as U, and (2) oxidation of Am to its
higher oxidation state. The CV and DPV studies revealed redox speciation of U matrices and
Am-oxide dissolved in DESs with insights into their redox thermodynamics, the trend
spectra of U(V) and U(IV) species in DESs. The differences in the redox parameters of
uranium species in DESs versus aqueous systems were elucidated in terms of the differences
EXAFS, which provided its coordination modes, stability (log K) and energetic of
interaction. The simulation studies (MD and DFT) corroborated the experimental results of
CV and DFT of dissolved uranium species in DESs and gave an insight into the optimized
geometry, binding energy, and MO diagram of U species in DESs. The present study enables
the analytical determination of U using DES by calibration plot. The simplicity of proposed
analysis methodology, short analysis time, high speed of sample preparation, lack of usage of
hazardous concentrated acids and oxidizing agents, and the use of safe and inexpensive
components recommends the high potential of Maline for routine analysis of U related to
eutectic solvent such as Malonic acid:ChCl (1:1) than the other DES with the stoichiometric
ratio of 1:2 or 2:1 (of HBD and HBA) for dissolution of metal oxides would be more
economical in terms of use of chemical/reagent for its practical application on large scale.
The findings of the present studies related to dissolution, speciation and analysis of U and
240
Am are of great significance in nuclear chemistry from both fundamental and application
point of view and show substantial fundamental advancement in the research area of actinide-
DES system. The knowledge gained by studies on speciation of U and Am solid samples in
the present DESs are significant for advancement of room temperature solvent media as an
hazardous acids and alkali metal chloride melts in high-temperature pyrochemical process.
dissertation are (i) generation of redox energetic, transport property and electron transfer
kinetic parameters of uranyl and neptunyl ions in presence of carboxylate and phosphonate
aqueous solution, (ii) assessing the impact of carboxylate and phosphonate based chelators on
the redox chemistry and oxidation state stabilization of uranium and neptunium, (iii)
providing stability constant (logβ) values of uranyl and neptunyl ions with above chelating
speciation of actinides in aqueous and deep eutectic solvent media, (v) discovering non-
aqueous Maline and imidazolium cation-malonic acid/diglycolic acid deep eutectic solvents
as a potential metal ion hijacker for processing diverse kind of solid uranium matrices
including oxides, salts, alloy and metal powder, and (vi) developing new greener and
sustainable methodology for direct sequestering and analysis of uranium in nuclear material
samples.
In this thesis, the problem of usage of hazardous concentrated inorganic acid during
processing of actinide matrices was addressed and it was suggested that environmental
benign DES with unique solvating potential can be a promising media to dissolve solid
matrices of actinides. The key interesting findings of present thesis can be described as (i)
241
existence of new pentavalent uranium species with citrate and phosphonocarboxylate
chelators in aqueous solution, (ii) the logβ value of pentavalent uranium with PC chelators is
intramolecular hydrogen bond (IMHB) with axial oxygen of uranyl, (iv) coupling of chemical
reaction with electron transfer step during reduction of uranyl in presence of citrate and
phosphonocarboxylates, and (v) choline dihydrogen citrate-malonic acid (1:1) DES works as
a promising electrolyte for dissolution of Am-oxide and selective oxidation of Am(III) to its
The works discussed in the present dissertation will potentially help in the design and
manipulation of redox-based remediation strategies of actinyl ions using the provided redox
evaluated results in the present studies would help in assessing strategies for (i) microbial
bioremediation of actinyl ions, and (ii) nuclear waste treatment and risk assessment for waste
disposal containing these ions. The findings of present research works have implications for
understanding the mechanism of biotic and abiotic reduction of actinyl species and their
biogeochemical cycles in a natural environment, and elucidating the pathways for the uptake
predict migration pathways of actinyl ions in geosphere upon their release from DGR. This
work makes a significant contribution to the fundamental understanding and prediction of the
Limitations of the present thesis can be described as below: The speciation studies in the
present thesis were performed at room temperature. The information about the redox and
molecular speciation of uranyl and neptunyl with varying temperature in aqueous and DES
media is beyond the scope the thesis. Herein, the non-aqueous solution was chosen as DES
242
and the studies about the speciation of U, Np and Am in other non-aqueous media such as
ionic liquids and organic solvents are not covered. The present studies involve the speciation
studies of actinides U and Np in aqueous solution while U and Am in DES media whereas the
speciation studies of actinides other than above in aqueous and DES media are outside the
scope of thesis. As bifunctional chelator, the PCs are selected for speciation studies of uranyl.
Thus, speciation studies of uranyl with other bifunctional chelator such as diphosphonates or
dicarboxylates are outside the context of present works. In the present studies, the DESs were
chosen only carboxylate-based moiety as HBD such as malonic acid and diglycolic acid. The
aim of present studies is speciation of actinyl ions therefore the only aqueous speciation
parameters such stoichiometry, stability, and structure of pentavalent species with citrate and
phosphonocarboxylates were reported and the isolation of pentavalent species is beyond the
Future outlook: Following are some suggested future research works based on the presented
works in this thesis: (i) Development of speciation sensor, (ii) Isolation of pentavalent species
redox speciation of actinides in aqueous and DES, (iv) Redox speciation of actinides at
modified electrode using ligand functionalized electrodic materials, (v) Aqueous speciation
(vi) Future design and synthesis of U selective novel adsorbents and molecular recognition of
actinides through activation of oxo group of actinyl using PCs based chelator (v)
Development of redox and non-redox based selective sequestration strategies of actinyl ions
using DESs and exploring hydrophobic DESs for their potential application in actinide
chemistry.
243
Appendix
Table S1.1. The reactions and parameters used in the aquatic speciation modelling
244
Chapter 2
Table S2.1. The amount of sample taken and their respective concentration in Maline.
U3Si2 17 2 8.5
Table S2.2. The amount of sample taken and their respective concentration of uranium
DES 2
U 3 O8 4.8
245
Appendix: Chapter 3
Figure S3.1. Speciation plots: (a) At [UO22+]= 10-5 M in presence of [Citrate]= 5x10-3 M; (b)
At [UO22+]= 10-3 M in presence of [Citrate]= 5x10-2 M; (c) At [UO22+]= 10-3 M in presence of
[Citrate]= 1x10-3 M.
246
-MS, 0.2-0.2min #(10-13)
x104
459.0829
406.0836
2
339.2299 919.1733
555.1064 619.0872
0
200 400 600 800 1000 1200 m/z
(a)
459.0640
1.0
0.8
0.6
351.5025 704.0155
439.0505
0.4 598.0182 725.9991
0.2 406.0622 675.0595
549.0519
829.5730 879.1030 919.1347
0.0
300 400 500 600 700 800 900 m/z
(b)
247
919.1
50
40
1018.9
30
20 568.5
1110.0
1020.2
566.6
1109.6
1119.6
10 807.5
1025.9
951.3 1157.9
784.0 1031.9
(c)
248
919.1
30
25
20
1018.9
15
568.5
1109.0
1027.7
660.1
1094.0
10
1110.3
190.9 980.8
674.6 1150.2
1026.1
5 945.3
558.3 1091.5
428.4 879.7 1187.5
(d)
249
919.1
60
50
459
40
30
981.8
20 1091.1
709.4
919.9
818.5 982.4
745.6
190.9 805.8 928.6
10 455.6 679.4 1035.8
Figure S3.2. ESI-MS data conditions: (a) [UO22+]= 10-5 M in presence of [Citrate]= 5x10-3 M
at pH 4, (b) Same as (a) at pH 6.5, (c) [UO22+]= 10-3 M in presence of [Citrate]= 5x10-2 M at
pH 4, (d) Same as (c) at pH 6.4, (e) [UO22+]= 10-3 M in presence of [Citrate]= 1x10-3 M at
pH 3.25, (f) & (g) Theoretically calculated m/z value of isotopic peak for [(UO2)2Cit2]2− and
[UO2Cit]− respectively, (h) Experimentally observed isotopic peak pattern.
250
Figure S3.3. Speciation plot of UO22+(10-5 M) in presence of 5x10-3 M PPA.
(a)
251
(b)
Figure S3.4. (a) & (b) are ESI--MS results at UO22+ (10-5M) and PPA (10-4M)
Appendix: Chapter-4
Table S4.1. Variation of ∆Ep and ipc/ipa with Scan rates for Np(VI)-PPA reduction (Peak I).
Table S4.2. Parameters obtained from Ep vs. logν plot for Np(V)-Phenylphosphonate
Phenylphosphonate (Peak
II)
Slope -0.054
αn 1.091
n (Taken α=0.5) 2.1
k0 (S-1) 0.258
252
Appendix: Chapter 5
Intens.
x104
437.2154
0.8
431.0269
0.6
403.8867 409.1905
447.9002
0.4
376.3275 467.0536
387.0395
0.2 381.0299
423.0541
393.3156
459.0801 482.8567
463.8603
0.0
380 400 420 440 460 480
-6 -5
(i) ESI-MS spectrum of uranyl(VI) (4x10 M) with PFA (4x10 M) in positive ion mode.
Intens.
x104
409.0232
1.25
485.0104
1.00
359.0306 499.0257
427.0387
445.0502
0.75
352.9783
470.9914
364.9347
506.9952 563.0500
0.50 433.1063 492.9774
513.0429 549.0276
455.0452
374.9626
571.0148
478.9608
0.25 536.9576
419.0897
400.0473
0.00
350 400 450 500 550
-6 -5
(ii) ESI-MS spectrum of uranyl(VI) (4x10 M) with PAA (4x10 M) in positive ion mode.
Intens.
x105
337.0539
359.0331
527.0683
2
375.0002 505.0845
392.9742
513.0481
491.0658
0
325 350 375 400 425 450 475 500 525
(iii) ESI-MS spectrum of uranyl(VI) (4x10-6M) with 3-PPA (4x10-5M) in positive ion mode.
Figure S5.1: (i-iii) ESI-MS spectra of Uranyl(VI) with PFA, PAA and 3-PPA respectively.
253
(a)
(b)
254
Table S5.1: The λmax with corresponding ε values for free UO22+ and its complexes with CPs
Table S5.2: The partial charges on the key atoms in the bare ligand and its corresponding
uranyl complex for all the three Uranyl(VI)-CP complexes.
255
Table S5.3: The O-O distances in bare and complexes forms and the subtended angles at
uranium by axial oxygens of Uranyl(VI) and the coordinating atoms of the ligand in the
complex for the Uranyl(VI)-CPs.
∟O-U-O of
Bare ligand Complex ∟O-U-O of Complex
System UO2 (in
(in A0) (in A0) (in degrees)
degrees)
256
Appendix: Chapter 6:
Maline
257
Scheme S6.2. Proposed schematic solvation structure model of UO22+ species in (a) Aqueous
258
Scheme S6.3. (a) Steps and chemicals/Reagents
chemicals/Reagents used in the previous method of hazardous
259
Scheme S6.3. (b) Reaction scheme and analysis process in the present developed
supramolecular scaffold.
260
(a) (b)
(c)
Figure S6.1. DPV plots (a) For the of UO3, UN, UO2, U3O8 and blank Maline and (b) For the
U std., U metal, U3Si2 and blank Maline (c) PuO2 and (U, Pu)O2.
261
80 12
g(r)U-Cl
60 9
water
g(r)U-O
g(r)U-Cl
40 6
20 3
0
0 2 4 6 8 10
0
0 2 4 6 8 10
o
r(A ) o
r(A )
a
(a) (b)
Figure S6.2. (a) and (b) are the calculated radial distribution function (RDF) for U-Cl- and
U-Owater.
90 12
75
60 8
mal
g(r)U-Br
g(r)U-O
45
30 4
15
0 0
0 2 4 6 8 10 0 2 4 6 8 10
o 0
r(A ) r(A )
(a) (b)
Figure S6.3. Calculated radial distribution function (RDF) for (a) U-Br- (b) U-Omalonic.
262
Table S6.1. The observed and calculated weight loss along with the predicted molecules being liberated at various temperatures in TG-
DSC. (In blank Maline and Uranium matrices dissolved Maline).
Name
of % Weight loss % Weight loss % Weight loss
Temperat Predicted Temperat Predicted Temperat Predicted
liquid
ure Molecules/at ure Molecules/at ure Molecules/
Regions of loss Regions of loss Regions of loss
Range of oms loss Range of oms loss Range of atoms loss
R1 R2 R3
R1 loss for from the loss R2 loss for from the loss R3 loss for from the loss
( ( (
(obser R1/ 0C in weight (observ R2/0C in weight (observ R3/0C in weight
Calculat Calculat Calculat
ved) ed) ed)
ed) ed) ed)
Maline
Choline
DES 43.3 43 100-266 Malonic acid - - 54.5 57.2 266-332
Chloride
Solutio
n of 2Choline 2malonate+8
44.5 42 100-242 8.33 9.0 242-276 2H2O+2H2O 44.1 276-320
UO3 in moity+7H2O 44.8 H2O
DES
Solutio
n of 2Choline 2malonate+7
40.5 42 100-240 15.2 13.7 242-286 2HCl+2H2O 41.8 41.8 286-326
UO2 in moity+7H2O H2O
DES
Solutio
n of 2Choline 2malonate+6
45.5 42 100-240 14.6 13.7 240-287 2HCl+2H2O 38.14 39.6 287-326
UN in moity+7H2O H2O
DES
263
Appendix: Chapter 6 part (b)
264
References
1. Grimes, Robin W., and William J. Nuttall. Science 329, no. 5993 (2010): 799-803.
2. Kharecha, Pushker A., and James E. Hansen. Environmental science & technology 47, no.
9 (2013): 4889-4895.
3. https://world-nuclear.org/information-library/current-and-future-generation/nuclear-power-
in-the-world-today.aspx
4. INTERNATIONAL ATOMIC ENERGY AGENCY, Nuclear Power Reactors in the
World, Reference Data Series No. 2, IAEA, Vienna (2021)
5. INTERNATIONAL ATOMIC ENERGY AGENCY, Climate Change and Nuclear Power
2018, Non-serial Publications, IAEA, Vienna (2018)
6. Frost, B. R. T. Royal Institute of Chemistry, Reviews 2, no. 2 (1969): 163-205.
7. The nuclear fuel cycle by IAEA Division of Nuclear Fuel Cycle and Waste
Technology, August 2011, 11-2522
8. OECD/NEA (2011), Trends towards Sustainability in the Nuclear Fuel Cycle, Nuclear
Development, OECD Publishing, Paris, https://doi.org/10.1787/9789264168268-en.
9. H. J. Bhabha, N. B. Prasad, In: Proceedings of 2nd UN International Conference on
‘Peaceful Uses of Atomic Energy’ Geneva, September 1-13, 1958, vol. 1, pp. 89-101.
10. INTERNATIONAL ATOMIC ENERGY AGENCY, Classification of Radioactive
Waste, IAEA Safety Standards Series No. GSG-1, IAEA, Vienna (2009)
11. Raj, Kanwar, K. K. Prasad, and N. K. Bansal. Nuclear Engineering and Design 236, no.
7-8 (2006): 914-930.
12. Pusch, R. Backfilling with mixture of bentonite/ballast materials or natural smectite clayǁ.
Swedish Nuclear Fuel and Waste management Co. (SKB), Stockholm. (1998), Technical
report, Tr-98-16.
13. Choppin, Gregory R. Radiochimica acta 32, no. 1-3 (1983): 43-54.
14. CHOPPIN, GREGORY R., and LIN F. RAO. Radiochimica Acta 37, no. 3 (1984): 143-
146.
15. Edelstein, Norman M., Jean Fuger, Joseph J. Katz, and Lester R. Morss. The chemistry of
the Actinide and Transactinide Elements (2006): 1753-1835.
16. Silva, R. J., and H. Nitsche Radiochimca Acta 70, no. Supplement (1995): 377-396.
17. Choppin, Gregory R. Radiochimica Acta 91, no. 11 (2003): 645-650.
18. Cook, Orator F. Science 23, no. 587 (1906): 506-507.
265
19. Altmaier, Marcus, Xavier Gaona, and Thomas Fanghänel. Chemical reviews 113, no. 2
(2013): 901-943.
20. Maher, Kate, John R. Bargar, and Gordon E. Brown Jr. Inorganic chemistry 52, no. 7
(2013): 3510-3532.
21. Jones, Matthew B., and Andrew J. Gaunt. Chemical Reviews 113, no. 2 (2013): 1137-
1198.
22. Bjørklund, Geir, Olav Albert Christophersen, Salvatore Chirumbolo, Olle Selinus, and
Jan Aaseth. Environmental research 156 (2017): 526-533.
23. Runde, W. H. Los Alamos Science 26 (2000): 392-411.
24. O’Loughlin, Edward J., Maxim I. Boyanov, Dionysios A. Antonopoulos, and Kenneth M.
Kemner. In Aquatic Redox Chemistry, pp. 477-517. American Chemical Society, 2011.
25. Rittmann, B. E., D. T. Reed, S. B. Aase, and A. J. Kropf. In AIP Conference Proceedings,
vol. 532, no. 1, pp. 63-64. American Institute of Physics, 2000.
26. Choppin, Gregory R. Radiochimica Acta 58, no. 1 (1992): 113-120.
27. Keith-Roach, Miranda J. Science of the Total Environment 396, no. 1 (2008): 1-11.
28. Szabó, Zoltan, Takashi Toraishi, Valerie Vallet, and Ingmar Grenthe. Coordination
Chemistry Reviews 250, no. 7-8 (2006): 784-815.
29. Choppin, Gregory R. Marine Chemistry 99, no. 1-4 (2006): 83-92.
30. Franczyk, Thaddeus S., Kenneth R. Czerwinski, and Kenneth N. Raymond. Journal of the
American Chemical Society 114, no. 21 (1992): 8138-8146.
31. Kannan, Shanmugaperumal, Mukesh Kumar, Biswajit Sadhu, Madhavan Jaccob, and
Mahesh Sundararajan. Dalton Transactions 46, no. 48 (2017): 16939-16946.
32. OECD-NEA. Physics and Safety of Transmutation Systems - A Status Report; Technical
Report NEA-6090; OECD: Paris, France, 2006
33. Fan, Fang-Li, Zhi Qin, Shi-Wei Cao, Cun-Min Tan, Qing-Gang Huang, De-Sheng Chen,
Jie-Ru Wang, Xiao-Jie Yin, Chao Xu, and Xiao-Gui Feng. Inorganic Chemistry 58, no. 1
(2018): 603-609.
34. INTERNATIONAL ATOMIC ENERGY AGENCY, Spent Fuel Reprocessing
Options, IAEA-TECDOC-CD-1587, IAEA, Vienna (2009).
35. Nash, Kenneth L., and Gregg J. Lumetta, eds. Elsevier, 2011.
36. INTERNATIONAL ATOMIC ENERGY AGENCY, IAEA Safeguards Glossary,
International Nuclear Verification Series No. 3 (Rev. 1), IAEA, Vienna (2022).
37. Meyers, Lisa A., Thomas M. Yoshida, Rebecca M. Chamberlin, and Ning Xu. Journal of
Radioanalytical and Nuclear Chemistry 310 (2016): 817-821.
266
38. Desigan, N., Nirav P. Bhatt, N. K. Pandey, U. Kamachi Mudali, R. Natarajan, and J. B.
Joshi. Journal of Radioanalytical and Nuclear Chemistry 312 (2017): 141-149.
39. Richter, Janine, and Michael Ruck. Molecules 25, no. 1 (2019): 78.
40. Wanigasekara, Eranda, John W. Freiderich, Xiao-Guang Sun, Roberta A. Meisner,
Huimin Luo, Lætitia H. Delmau, Sheng Dai, and Bruce A. Moyer. Dalton Transactions 45,
no. 25 (2016): 10151-10154.
41. Mohapatra, Prasanta Kumar. Dalton Transactions 46, no. 6 (2017): 1730-1747.
42. Sun, Xiaoqi, Huimin Luo, and Sheng Dai Chemical reviews 112, no. 4 (2012): 2100-
2128.
43. Mahanty, Bholanath, and Prasanta Kumar Mohapatra. Separation Science and
Technology 57, no. 17 (2022): 2792-2823.
44. Nockemann, Peter, Ben Thijs, Stijn Pittois, Jan Thoen, Christ Glorieux, Kristof Van
Hecke, Luc Van Meervelt, Barbara Kirchner, and Koen Binnemans. The Journal of Physical
Chemistry B 110, no. 42 (2006): 20978-20992.
45. Smith, Emma L., Andrew P. Abbott, and Karl S. Ryder. Chemical reviews 114, no. 21
(2014): 11060-11082.
46. Richter, Janine, and Michael Ruck. RSC advances 9, no. 51 (2019): 29699-29710.
47. Jones, Matthew B., and Andrew J. Gaunt. Chemical Reviews 113, no. 2 (2013): 1137-
1198.
48. Hartley, Jennifer M., Chung-Man Ip, Gregory CH Forrest, Kuldip Singh, Stephen J.
Gurman, Karl S. Ryder, Andrew P. Abbott, and Gero Frisch. Inorganic chemistry 53, no. 12
(2014): 6280-6288.
49. Venkatesan, K. A., Jagadeeswara Rao, K. Nagarajan, and P. R. Vasudeva
Rao. International Journal of Electrochemistry 2012 (2012).
50. Ramakumar, K. L. International Journal of Nuclear Safety and Security 1, no. 2 (2022):
96-103.
51. Mayer, K. In Seminar on Modern Verification Regimes: Similarities, Synergies and
Challenges, 12–14 May 1998, Helsinki, Finland; Report EUR 18681EN, pp. 189-196. 1998.
52. Nair, P., K. Lohithakshan, Mary Xavier, S. Marathe, and H. Jain. Journal of
Radioanalytical and Nuclear Chemistry 122, no. 1 (1988): 19-26.
53. Savosina, Julia, Marina Agafonova-Moroz, Irina Yaroshenko, Julia Ashina, Vasily
Babain, Alexander Lumpov, Andrey Legin, and Dmitry Kirsanov. Sensors 20, no. 6 (2020):
1604.
267
54. Bersier, Pierre M., Jonathon Howell, and Craig Bruntlett Analyst 119, no. 2 (1994): 219-
232.
55. Lohrengel, M. M., and J. W. Schultze Electrochimica Acta 21, no. 11 (1976): 957-965.
56. Brennan, M. P. J., and O. R. Brown. Journal of Applied Electrochemistry 2, no. 1 (1972):
43-49.
57. Shishov, Andrey, Andrey Bulatov, Marcello Locatelli, Simone Carradori, and Vasil
Andruch. Microchemical journal 135 (2017): 33-38.
58. Selbin, Joel, and J. D. Ortego Chemical Reviews 69, no. 5 (1969): 657-671.
59. Burdet, Fabien, Jacques Pécaut, and Marinella Mazzanti. Journal of the American
Chemical Society 128, no. 51 (2006): 16512-16513.
60. Nocton, Gregory, Pawel Horeglad, Jacques Pécaut, and Marinella Mazzanti. Journal of
the American Chemical Society 130, no. 49 (2008): 16633-16645.
61. Donnet, L., J. M. Adnet, N. Faure, P. Bros, Ph Brossard, and F. Josso. In Proceedings of
the 5th OECD-NEA International Information Exchange Meeting on Actinide and Fission
Product Partitioning and Transmutation–Session II (Partitioning). SCK-CEN, Mol, Belgium,
Nov, pp. 25-27. 1998.
62. Abbott, Andrew P., Glen Capper, David L. Davies, Helen L. Munro, Raymond K.
Rasheed, and Vasuki Tambyrajah. Chemical Communications 19 (2001): 2010-2011.
63. Abbott, Andrew P., Glen Capper, David L. Davies, Raymond K. Rasheed, and Vasuki
Tambyrajah. Chemical communications 1 (2003): 70-71.
64. Abbott, Andrew P., David Boothby, Glen Capper, David L. Davies, and Raymond K.
Rasheed. Journal of the American Chemical Society 126, no. 29 (2004): 9142-9147.
65. Martins, Mónia AR, Simão P. Pinho, and João AP Coutinho. Journal of Solution
Chemistry 48 (2019): 962-982.
66. El Achkar, Tracy, Hélène Greige-Gerges, and Sophie Fourmentin Environmental
chemistry letters 19 (2021): 3397-3408.
67. Zhang, Qinghua, Karine De Oliveira Vigier, Sebastien Royer, and François Jérôme.
Chemical Society Reviews 41, no. 21 (2012): 7108-7146.
68. Mutlu, Hatice, and Leonie Barner. Macromolecular Chemistry and Physics 223, no. 13
(2022): 2200111.
69. Hutzinger, Otto. Environmental Science and Pollution Research International 6, no. 3
(1999): 123.
70. Billard, Isabelle, Clotilde Gaillard, and Christoph Hennig. Dalton Transactions 37
(2007): 4214-4221.
268
71. Colombo Dugoni, Greta, Eros Mossini, Elena Macerata, Alessandro Sacchetti, Andrea
Mele, and Mario Mariani. ACS omega 6, no. 5 (2021): 3602-3611.
72. Ikeda-Ohno, Atsushi, Christoph Hennig, André Rossberg, Harald Funke, Andreas C.
Scheinost, Gert Bernhard, and Tsuyoshi Yaita. Inorganic chemistry 47, no. 18 (2008): 8294-
8305.
73. McNaught, Alan D. Vol. 1669. Oxford: Blackwell Science, 1997.
74. Varma, Rajender S., and Vasudevan V. Namboodiri. Pure and Applied Chemistry 73, no.
8 (2001): 1309-1313.
75. A.J. Bard, L.R. Faulkner, Electrochemical Methods—Fundamentals and Applications,
Wiley, New York, 1980
76. Zanello, P., Nervi, C. and De Biani, F.F., Inorganic electrochemistry: theory, practice and
application. Royal Society of Chemistry, 2019.
77. Analytical Electrochemistry, Second Edition. Joseph Wang Wiley-VCH
ISBNs: 0-471-28272-3
78. Kuwana, Theodore, R. K. Darlington, and D. W. Leedy. Analytical Chemistry 36, no. 10
(1964): 2023-2025.
79. Ciglenečki, I., M. Marguš, and P. Orlović-Leko. Int J Biosen Bioelectron 4, no. 3 (2018):
94-96.
80. Gans, Peter, Antonio Sabatini, and Alberto Vacca. Talanta 43, no. 10 (1996): 1739-1753.
81. Poswal, A. K., A. Agrawal, A. K. Yadav, C. Nayak, S. Basu, S. R. Kane, C. K. Garg, D.
Bhattachryya, S. N. Jha, and N. K. Sahoo In AIP Conference Proceedings, vol. 1591, no. 1,
pp. 649-651. American Institute of Physics, 2014.
82. Newville, M., B. Ravel, D. Haskel, J. J. Rehr, E. A. Stern, and Y(1995 Yacoby. Physica
B: Condensed Matter 208 (1995): 154-156.
83. Banerjee, Shibdas, and Shyamalava Mazumdar. International journal of analytical
chemistry 2012 (2012).
84. Zhao, Yan, and Donald G. Truhlar. Accounts of chemical research 41, no. 2 (2008): 157-
167
85. Schäfer, Ansgar, Christian Huber, and Reinhart Ahlrichs. The Journal of chemical
physics 100, no. 8 (1994): 5829-5835.
86. Ahlrichs, Reinhart, Michael Bär, Marco Häser, Hans Horn, and Christoph
Kölmel. Chemical Physics Letters 162, no. 3 (1989): 165-169.
87. Martínez, Leandro, Ricardo Andrade, Ernesto G. Birgin, and José Mario
Martínez. Journal of computational chemistry 30, no. 13 (2009): 2157-2164.
269
88. Lindahl, Erik, Berk Hess, and David Van Der Spoel. Molecular modeling annual 7
(2001): 306-317.
89. Sahu, Pooja, Sk Musharaf Ali, and Kalasanka Trivikram Shenoy. Physical Chemistry
Chemical Physics 18, no. 34 (2016): 23769-23784.
90. Alderighi, Lucia, Peter Gans, Andrea Ienco, Daniel Peters, Antonio Sabatini, and Alberto
Vacca. Coordination chemistry reviews 184, no. 1 (1999): 311-318.
91. Herndon, Virginia. "MINTEQA2/PRODEFA2, A Geochemical Assessment model for
Environmental Systems: User Manual Supplement for Version 4.0." (1998).
92. Velasco, Jaime González. Electroanalysis 9, no. 11 (1997): 880-882.
93. DeFord, Donald D., and David N. Hume. Journal of the American Chemical Society 73,
no. 11 (1951): 5321-5322.
94. J. S. Renny, L. L. Tomasevich, E. H. Tallmadge, D. B. Collum, Ang. Chem. Int. Ed.
Eng. 46 (2013) 11998 – 12013.
95. Skoog, D. A., D. M. West, and F. J. Holler. "Titrimetric Methods of
analysis." Fundamentals of Analytical Chemistry. 5th ed., p100-121, Saunders, New
York (1988).
96. Dicks, Andrew P., and Andrei Hent. Green chemistry metrics: a guide to determining and
evaluating process greenness. London: Springer International Publishing, 2015.
97. Berto, Silvia, Francesco Crea, Pier Giuseppe Daniele, Antonio Gianguzza, Alberto
Pettignano, and Silvio SammartanoCoordination Chemistry Reviews 256, no. 1-2 (2012): 63-
81.
98. Morris, David E. Inorganic Chemistry 41, no. 13 (2002): 3542-3547.
99. Bailey, E. H., J. F. W. Mosselmans, and P. F. Schofield. Chemical Geology 216, no. 1-2
(2005): 1-16.
100. Dodge, Cleveland J., and Arokiasamy J. Francis. Environmental science &
technology 28, no. 7 (1994): 1300-1306.
101. Francis, A. J., C. J. Dodge, and J. B. Gillow. Nature 356, no. 6365 (1992): 140-142.
102. Basile, M., D. K. Unruh, K. Gojdas, E. Flores, L. Streicher, and T. Z. Forbes. Chemical
Communications 51, no. 25 (2015): 5306-5309.
103. Clearfield, Abraham. Current Opinion in Solid State and Materials Science 1, no. 2
(1996): 268-278.
104. Clearfield, Abraham. Current Opinion in Solid State and Materials Science 6, no. 6
(2002): 495-506.
270
105. Barančíková, Gabriela, Tibor Liptaj, and Nadežda Prónayová. Soil and water
research 2, no. 4 (2007): 141.
106. Nash, K. L. Journal of alloys and compounds 213 (1994): 300-304.
107. Stone, A. T., M. A. Knight, and B. Nowack. 2002.
108. Nash, Κ. L. Radiochimica Acta 61, no. 3-4 (1993): 147-154.
109. Jerden Jr, James L., and A. K. Sinha. Journal of Geochemical Exploration 91, no. 1-3
(2006): 56-70.
110. Feldman, Isaac, and W. F. Neuman. Journal of the American Chemical Society 73, no. 5
(1951): 2312-2315.
111. Rajan, K. S., and Arthur Earl Martell. Inorganic Chemistry 4, no. 4 (1965): 462-469.
112. Pasilis, Sofie P., and Jeanne E. Pemberton. Inorganic chemistry 42, no. 21 (2003): 6793-
6800.
113. Berto, Silvia, Francesco Crea, Pier G. Daniele, Concetta De Stefano, Enrico Prenesti,
and Silvio Sammartano. (2012): 13-28.
114. Basile, Madeline, Daniel K. Unruh, Erin Flores, Adam Johns, and Tori Z. Forbes.
Dalton Transactions 44, no. 6 (2015): 2597-2605.
115. Suzuki, Yoshinori, T. Nankawa, T. Yoshida, Takuo Ozaki, T. Ohnuki, Arokiasamy J.
Francis, S. Tsushima, Y. Enokida, and I. Yamamoto. Radiochimica Acta 94, no. 9-11 (2006):
579-583.
116. Harris, Walter Edgar, and I. M. Kolthoff. Journal of the American Chemical Society 67,
no. 9 (1945): 1484-1490.
117. Kolthoff, I. M., and W. E. Harris. Journal of the American Chemical Society 68, no. 7
(1946): 1175-1179.
118. Harris, W. E., and I. M. Kolthoff. Journal of the American Chemical Society 69, no. 2
(1947): 446-451.
119. Suzuki, Yoshinori, Takuya Nankawa, Takuo Ozaki, Toshihiko Ohnuki, Arokiasamy J.
Francis, Youichi Enokida, and Ichiro Yamamoto. Journal of nuclear science and
technology 44, no. 9 (2007): 1227-1232.
120. Hardwick, Helen C., Drew S. Royal, Madeleine Helliwell, Simon JA Pope, Lorna
Ashton, Roy Goodacre, and Clint A. Sharrad Dalton Transactions 40, no. 22 (2011): 5939-
5952.
121. Randles, John EB. Transactions of the Faraday Society 44 (1948): 327-338.
122. Ševčík, A. Collection of Czechoslovak Chemical Communications 13 (1948): 349-377.
271
123. P. Zanello, Inorganic Electrochemistry: Theory Practice and Applications, Royal Society
of Chemistry, Cambridge, UK, 2003, pp. 61.
124. Baker, Blair C., and Donald T. Sawyer. Inorganic Chemistry 9, no. 2 (1970): 197-204.
125. R. S. Nicholson, I. Shain, Anal. Chem., 1964, 36, 706-723.
126. R. S. Nicholson, I. Shain, Anal. Chem., 1965, 37, 178-190.
127. C. Musikas, J. inorg. nucl. Chem., Supplement 1976. Pergamon Press., Printed in Great
Britain.
128. P. Kissinger, W. R. Heineman, Laboratory techniques in electroanalytical chemistry,
2nd ed. Marcel Dekker, INC: New York, N.Y., 1996, pp-60.
129. Mizuoka, Koichiro, Seong-Yun Kim, Miki Hasegawa, Toshihiko Hoshi, Gunzo
Uchiyama, and Yasuhisa Ikeda. Inorganic chemistry 42, no. 4 (2003): 1031-1038.
130. Mizuguchi, Kohji, Yoon-Yul Park, Hiroshi Tomiyasu, and Yasuhisa Ikeda. Journal of
Nuclear Science and Technology 30, no. 6 (1993): 542-548.
131. Mizuoka, Koichiro, Satoru Tsushima, Miki Hasegawa, Toshihiko Hoshi, and Yasuhisa
Ikeda. Inorganic chemistry 44, no. 18 (2005): 6211-6218.
132. Cohen, Donald Journal of Inorganic and Nuclear Chemistry 32, no. 11 (1970): 3525-
3530.
133. Bell, J. T., H. A. Friedman, and M. R. Billings. Journal of Inorganic and Nuclear
Chemistry 36, no. 11 (1974): 2563-2567.
134. Galindo, Catherine, Mirella Del Nero, Remi Barillon, Eric Halter, and Benoit Made.
Journal of colloid and interface science 347, no. 2 (2010): 282-289.
135. Kraus, Kurt A., and Frederick Nelson. Journal of the American Chemical Society 71, no.
7 (1949): 2517-2522.136. Kllngler, R. J., Kochl, J. K.: Electron-Transfer Kinetics from
Cyclic Voltammetry. Quantitative Description of Electrochemical Reversibility. J. Phys.
Chem. 85, 1731 (1981).
136. Kllngler, R. J., Kochl, J. K. J. Phys. Chem. 85, 1731 (1981).
137. Ansari, Seraj A., Arunasis Bhattacharyya, Zhicheng Zhang, and Linfeng Rao Inorganic
Chemistry 54, no. 17 (2015): 8693-8698.
138. Z. Yoshida, S.G. Johnson, T. Kimura, J.R. Krsul, Neptunium, Springer, Heidelberg,
Germany, 2005, p. 669.
139. S. Fried, Neptunium, McGraw-Hill, New York, 1954, p. 471.
140. Viswanathan, Hari S., Bruce A. Robinson, Albert J. Valocchi, and Ines R. Triay. Journal
of Hydrology 209, no. 1-4 (1998): 251-280.
141. McMillan, Edwin, and Philip Hauge Abelson. Physical Review 57, no. 12 (1940): 1185.
272
142. J.J. Katz, G.T. Seaborg, L.R. Morss, The chemistry of the actinide elements, Chapman
Hall: Lond. 1 (1986) 469–487.
143. Shcherbina, Natalia S., Irina V. Perminova, Stepan N. Kalmykov, Anton N. Kovalenko,
Richard G. Haire, and Alexander P. Novikov. Environmental science & technology 41, no. 20
(2007): 7010-7015.
144. Chatterjee, Sayandev, Samuel A. Bryan, Amanda J. Casella, James M. Peterson, and
Tatiana G. Levitskaia. Inorganic Chemistry Frontiers 4, no. 4 (2017): 581-594.
145. Ikeda-Ohno, Atsushi, Satoru Tsushima, Koichiro Takao, André Rossberg, Harald Funke,
Andreas C. Scheinost, Gert Bernhard, Tsuyoshi Yaita, and Christoph Hennig. Inorganic
chemistry 48, no. 24 (2009): 11779-11787.
146. Hennig, Christoph, Atsushi Ikeda-Ohno, Satoru Tsushima, and Andreas C. Scheinost
Inorganic chemistry 48, no. 12 (2009): 5350-5360.
147. M.P. Jensen, J.V. Beitz, R.D. Rogers, K.L. Nash, Dalt. Trans. 18 (2000) 3058–3064.
148. A. Srivastava, P. Kumar, B.S. Tomar, Radiochim. Acta 105 (2017) 311–320.
149. Nash, K. L. J. Alloys Cmpd 249 (1997): 33.
150. Takao, Koichiro, Shinobu Takao, Andreas C. Scheinost, Gert Bernhard, and Christoph
Hennig. Inorganic chemistry 48, no. 18 (2009): 8803-8810.
151. L. Rao, G. Tian, Complexation of actinid es with derivatives of oxydiacetic acid. In
Actinides 2005, Basic Science, Applications and Technology: Proceedings of the Materials
Research Society Symposium; (Eds. J. L. Sarrao, A. J. Schwartz) MRS: Warrendale, PA,
2006; Vol. 893; Paper 0893-JJ08-06.
152. Rao, Linfeng, Guoxin Tian, and Simon J. Teat. Dalton Transactions 39, no. 13 (2010):
3326-3330.
153. May, Iain, Robin J. Taylor, Iain S. Denniss, Geoff Brown, Andrew L. Wallwork, Nick J.
Hill, Jeremy M. Rawson, and Robert Less. Journal of alloys and compounds 275 (1998):
769-772.
154. B. J. Colston, G. R. Choppin, R. J. Taylor, Radiochim. Acta. 88 (2000) 329−334.
155. Sinkov, Sergei I., Gregory R. Choppin, and Robin J. Taylor. In ABSTRACTS OF
PAPERS OF THE AMERICAN CHEMICAL SOCIETY, vol. 232, pp. 853-853. 1155 16TH
ST, NW, WASHINGTON, DC 20036 USA: AMER CHEMICAL SOC, 2006.
156. Bray, Travis H., Anna-Gay D. Nelson, Geng Bang Jin, Richard G. Haire, and Thomas E.
Albrecht-Schmitt. Inorganic chemistry 46, no. 26 (2007): 10959-10961.
157. Nelson, Anna-Gay D., Travis H. Bray, Wei Zhan, Richard G. Haire, Todd S. Sayler, and
Thomas E. Albrecht-Schmitt. Inorganic chemistry 47, no. 11 (2008): 4945-4951.
273
158. Kim, S-Y., T. Asakura, Y. Morita, G. Uchiyama, and Y. Ikeda. Journal of
radioanalytical and nuclear chemistry 262, no. 2 (2004): 311-315.
159. Kitatsuji, Yoshihiro, Takaumi Kimura, and Sorin Kihara. Journal of Electroanalytical
Chemistry 641, no. 1-2 (2010): 83-89.
160. Niese, U., and J. Vecernik. (1982): 191-192.
161. E. Laviron, J Electroanal. Chem. 101 (1979) 19.
162. Yang, Yanqiu, Zhicheng Zhang, Guokui Liu, Shunzhong Luo, and Linfeng Rao The
Journal of Chemical Thermodynamics 80 (2015): 73-78.
163. Allen, P. G., J. J. Bucher, D. K. Shuh, N. M. Edelstein, and T. Reich. Inorganic
chemistry 36, no. 21 (1997): 4676-4683.
164. Rama Mohana Rao Dumpala, N. Rawat, A. Bhattacharya, B. S. Tomar, ChemSlct. 2
(2017) 2722-2731.
165. Srivastava, Ashutosh, Rama Mohan Rao Dumpala, Neetika Rawat, and B. S.
Tomar. Inorganica Chimica Acta 482 (2018): 307-316.
166. Craddock, Henry A. Oilfield chemistry and its environmental impact. John Wiley &
Sons, 2018.
167. Sullivan, P.J., O'Brian, F. and Ferguson, R., 1996, January. A Multifunctional Additive
for Deposit Control. In CORROSION 96. NACE International.
168. Mertens, Martha, Sebastian Höss, Günter Neumann, Joshua Afzal, and Wolfram
Reichenbecher. Environmental Science and Pollution Research 25 (2018): 5298-5317.
169. Kanissery, Ramdas, Biwek Gairhe, Davie Kadyampakeni, Ozgur Batuman, and
Fernando Alferez. Plants 8, no. 11 (2019): 499.
170. Weber, Jerzy, Yona Chen, Elżbieta Jamroz, and Teodoro Miano. Journal of Soils and
Sediments 18, no. 8 (2018): 2665-2667.
171. Koban, A., and G. Bernhard. Journal of inorganic biochemistry 101, no. 5 (2007): 750-
757.
172. Salome, Kathleen R., Melanie J. Beazley, Samuel M. Webb, Patricia A. Sobecky, and
Martial Taillefert Geochimica et Cosmochimica Acta 197 (2017): 27-42.
173. Knope, Karah E., and Christopher L. Cahill. Inorganic Chemistry Communications 13,
no. 9 (2010): 1040-1042.
174. Knope, Karah E., and Christopher L. Cahill. Inorganic chemistry 48, no. 14 (2009):
6845-6851.
175. Adelani, Pius O., and Thomas E. Albrecht-Schmitt. Crystal growth & design 11, no. 10
(2011): 4676-4683.
274
176. Wu, Dai, Xiaojing Bai, Hong-Rui Tian, Weiting Yang, Zewen Li, Qing Huang, Shiyu
Du, and Zhong-Ming Sun Inorganic Chemistry 54, no. 17 (2015): 8617-8624.
177. Görller-Walrand, C., and S. De Jaegere. Spectrochimica Acta Part A: Molecular
Spectroscopy 28, no. 2 (1972): 257-268.
178. Bell, J.T. and Biggers, R.E.,. Journal of Molecular Spectroscopy, 18(3), (1965).
179. Bliznyuk, V. N., N. A. Conroy, Y. Xie, R. Podila, A. M. Rao, and Brian A. Powell.
Physical Chemistry Chemical Physics 20, no. 3 (2018): 1752-1760.
180. Srivastava, Ashutosh, Ashis K. Satpati, Ritu Singh, Pranaw Kumar, Sumit Kumar, and
Bhupendra S. Tomar. New Journal of Chemistry 41, no. 24 (2017): 15094-15104.
181. Moll, H., G. Geipel, T. Reich, G. Bernhard, Th Fanghänel, and I. Grenthe. Radiochimica
Acta 91, no. 1 (2003): 11-20.
182. Starck, Matthieu, Fanny A. Laporte, Stephane Oros, Nathalie Sisommay, Vicky Gathu,
Pier Lorenzo Solari, Gaëlle Creff et al. Chemistry–A European Journal 23, no. 22 (2017):
5281-5290.
183. Barkleit, Astrid, Harald Foerstendorf, Bo Li, André Rossberg, Henry Moll, and Gert
Bernhard. Dalton transactions 40, no. 38 (2011): 9868-9876.
184. Capdevila, H., and P. Vitorge. Journal of radioanalytical and nuclear chemistry 143, no.
2 (1990): 403-414.
185. Ikeda, Atsushi, Christoph Hennig, Satoru Tsushima, Koichiro Takao, Yasuhisa Ikeda,
Andreas C. Scheinost, and Gert Bernhard. Inorganic chemistry 46, no. 10 (2007): 4212-4219.
186. Faizova, Radmila, Rosario Scopelliti, Anne-Sophie Chauvin, and Marinella Mazzanti.
Journal of the American Chemical Society 140, no. 42 (2018): 13554-13557.
187. Faizova, Radmila, Sarah White, Rosario Scopelliti, and Marinella Mazzanti Chemical
Science 9, no. 38 (2018): 7520-7527.
188. Fortier, Skye, and Trevor W. Hayton. Coordination Chemistry Reviews 254, no. 3-4
(2010): 197-214.
189. Lewis, Andrew J., Kimberly C. Mullane, Eiko Nakamaru-Ogiso, Patrick J. Carroll, and
Eric J. Schelter. Inorganic Chemistry 53, no. 13 (2014): 6944-6953.
190. Xiao, Chengliang, Zohreh Hassanzadeh Fard, Debajit Sarma, Tze-Bin Song, Chao Xu,
and Mercouri G. Kanatzidis Journal of the American Chemical Society 139, no. 46 (2017):
16494-16497.
191. Rodríguez-Penalonga, Laura, and B. Yolanda Moratilla Soria. Energies 10, no. 8 (2017):
1235.
275
192. Sun, Xiaoqi, Huimin Luo, and Sheng Dai. Chemical reviews 112, no. 4 (2012): 2100-
2128.
193. Lopez, Michael J., Matthew V. Sheridan, Jeffrey R. McLachlan, Travis S. Grimes, and
Christopher J. Dares. Chemical Communications 55, no. 28 (2019): 4035-4038.
194. Dares, Christopher J., Alexander M. Lapides, Bruce J. Mincher, and Thomas J. Meyer.
Science 350, no. 6261 (2015): 652-655.
195. Wellens, S., 2014. Ionic liquid technology in metal refining: dissolution of metal oxides
and separation by solvent extraction. KU Leuven, pp.1-100.
196. Abbott, Andrew P., Gero Frisch, Jennifer Hartley, and Karl S. Ryder. Green
Chemistry 13, no. 3 (2011): 471-481.
197. Xu, Qian, T. S. Zhao, Lei Wei, Cheng Zhang, and X. L. Zhou. Electrochimica Acta 154
(2015): 462-467.
198. Lloyd, David, Tuomas Vainikka, Lasse Murtomäki, Kyösti Kontturi, and Elisabet
Ahlberg Electrochimica Acta 56, no. 14 (2011): 4942-4948.
199. Brett, C.M., 2018. Current Opinion in Electrochemistry, 10, pp.143-148.
200. Abbott, Andrew P., Glen Capper, David L. Davies, Katy J. McKenzie, and Stephen U.
Obi. Journal of Chemical & Engineering Data 51, no. 4 (2006): 1280-1282.
201. Ünlü, Ayşe Ezgi, Azime Arıkaya, and Serpil Takaç. Green Processing and Synthesis 8,
no. 1 (2019): 355-372.
202. Söldner, Anika, Julia Zach, and Burkhard König. Green Chemistry 21, no. 2 (2019):
321-328.
203. Mukhopadhyay, Soumyadeep, Sumona Mukherjee, Nor Farihah Adnan, Adeeb Hayyan,
Maan Hayyan, Mohd Ali Hashim, and Bhaskar Sen Gupta. Chemical Engineering
Journal 294 (2016): 316-322.
204. Klähn, Marco, Abirami Seduraman, and Ping Wu. The Journal of Physical Chemistry
B 112, no. 44 (2008): 13849-13861.
205. Dupont, Jairton, and Paulo AZ Suarez. Physical Chemistry Chemical Physics 8, no. 21
(2006): 2441-2452.
206. Zhang, Chengdong. "Interactions of ionic liquids with uranium and its implications on
biotransformation." (2006).
207. S. Picart et al., ATALANTE-2004 P2-P4. Nimes, France, June 21-25 (2004).
208. Wilden, Andreas, Giuseppe Modolo, Peter Kaufholz, Fabian Sadowski, Steve Lange,
Michal Sypula, Daniel Magnusson, Udo Müllich, Andreas Geist, and Dirk Bosbach. Solvent
extraction and ion exchange 33, no. 2 (2015): 91-108.
276
209. Liu, Yang, J. Brent Friesen, James B. McAlpine, David C. Lankin, Shao-Nong Chen,
and Guido F. Pauli. Journal of natural products 81, no. 3 (2018): 679-690.
210. Delgado-Mellado, Noemí, Marcos Larriba, Pablo Navarro, Victoria Rigual, Miguel
Ayuso, Julián García, and Francisco Rodríguez. Journal of Molecular Liquids 260 (2018):
37-43.
211. Mishra, Satyabrata, K. Sini, Ch Jagadeeswara Rao, C. Mallika, and U. Kamachi Mudali.
Journal of Electroanalytical Chemistry 776 (2016): 127-133.
212. Srivastava, Ashutosh, Rama Mohana Rao Dumpala, Pranaw Kumar, Ravi Kumar, and
Neetika Rawat. Inorganic Chemistry 61, no. 39 (2022): 15452-15462.
213. Renjith, Anu, and V. Lakshminarayanan. The Journal of Physical Chemistry C 122, no.
44 (2018): 25411-25421.
214. Abbott, Andrew P., Glen Capper, and Stephen Gray Chemphyschem: a European
journal of chemical physics and physical chemistry 7, no. 4 (2006): 803-806.
215. Hennig, C., J. Tutschku, A. Rossberg, G. Bernhard, and A. C. Scheinost. Inorganic
chemistry 44, no. 19 (2005): 6655-6661.
216. Florindo, Catarina, Filipe S. Oliveira, Luis Paulo N. Rebelo, Ana M. Fernandes, and
Isabel M. Marrucho. ACS Sustainable Chemistry & Engineering 2, no. 10 (2014): 2416-2425.
217. Te Velde, G.T., Bickelhaupt, F.M., Baerends, E.J., Fonseca Guerra, C., van Gisbergen,
S.J., Snijders, J.G. and Ziegler, T., 2001. Chemistry with ADF. Journal of Computational
Chemistry, 22(9), pp.931-967.
218. ADF2017, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The
Netherlands, http://www.scm.com.
219. Perkins, Sasha L., Paul Painter, and Coray M. Colina. Journal of Chemical &
Engineering Data 59, no. 11 (2014): 3652-3662.
220. Anastas, P. and Eghbali, N., 2010. Green chemistry: principles and practice. Chemical
Society Reviews, 39(1), pp.301-312
221. Horváth, I.T., 2018. Introduction: Sustainable Chemistry. Chemical Reviews, 118(2),
pp.369-371.
222. Namieśnik, Jacek. Journal of separation Science 24, no. 2 (2001): 151-153.
223. Garcia, Gregorio, Santiago Aparicio, Ruh Ullah, and Mert Atilhan. Energy & Fuels 29,
no. 4 (2015): 2616-2644.
224. Aleksa, V. Lithuanian Journal of Physics and Technical Sciences 47, no. 4 (2007): 435-
441.
277
225. Zhao, Shu-Feng, Mike Horne, Alan M. Bond, and Jie Zhang. The Journal of Physical
Chemistry C 120, no. 42 (2016): 23989-24001.
226. Gorodetsky, Brian, Taramatee Ramnial, Neil R. Branda, and Jason AC
Clyburne. Chemical communications 17 (2004): 1972-1973.
227. Stonhill, L. G. Canadian Journal of Chemistry 36, no. 11 (1958): 1487-1492.
228. Nockemann, Peter, Ben Thijs, Tatjana N. Parac-Vogt, Kristof Van Hecke, Luc Van
Meervelt, Bernard Tinant, Ingo Hartenbach et al. Inorganic Chemistry 47, no. 21 (2008):
9987-9999.
229. Nockemann, Peter, Rik Van Deun, Ben Thijs, Diederik Huys, Evert Vanecht, Kristof
Van Hecke, Luc Van Meervelt, and Koen Binnemans. Inorganic chemistry 49, no. 7 (2010):
3351-3360.
230. Protsenko, V. S., L. S. Bobrova, A. A. Kityk, and F. I. Danilov. Journal of
Electroanalytical Chemistry 864 (2020): 114086.
231. Srivastava, Ashutosh, Rama Mohana Rao Dumpala, Pooja Sahu, Ashok Kumar Yadav,
Neetika Rawat, S. K. Musharaf Ali, Manjulata Sahu, Nimai Pathak, and Arijit Sengupta. ACS
Sustainable Chemistry & Engineering 9, no. 23 (2021): 7846-7862.
232. Kolbeck, C., J. Lehmann, K. R. J. Lovelock, T. Cremer, N. Paape, P. Wasserscheid, A.
P. Froba, F. Maier, and H-P. Steinruck. The Journal of Physical Chemistry B 114, no. 51
(2010): 17025-17036.
233. Srivastava, Ashutosh, Pooja Sahu, M. S. Murali, Sk Musharaf Ali, Manjulata Sahu, Jisha
S. Pillai, and Neetika Rawat. Journal of Electroanalytical Chemistry 901 (2021): 115752.
278
Publications
2. Redox speciation of uranyl in citrate medium: kinetics and reduction mechanism with in
situ spectroelectrochemical investigation. Ashutosh Srivastava et al. “New Journal of
Chemistry” 41, no. 24 (2017): 15094-15104.
4. New Greener and Sustainable Methodology for Direct Sequestering and Analysis of
Uranium Using a Maline Supramolecular Scaffold and Mechanistic Understanding through
Speciation and Interaction Studies. Ashutosh Srivastava et al. “ACS Sustainable Chem.
Eng.” 2021, 9, 23, 7846–7862.
5. New deep eutectic solvents based on imidazolium cation: Probing redox speciation of
uranium oxides by electrochemical and theoretical simulations. Ashutosh Srivastava et al.
“Journal of Electroanalytical Chemistry” 901 (2021): 115752.
1. Dumpala, Rama Mohana Rao, Ashutosh Srivastava, and Neetika Rawat. "Experimental
and theoretical approach to probe the aquatic speciation of transuranic (neptunyl) ion in
presence of two omnipresent organic moieties." Chemosphere 273 (2021): 129745.
2. Rao, Ankita, and Ashutosh Srivastava. "Supercritical carbon dioxide and eutectic solvent
in conjunction: Novel method for in-situ solvent preparation-dissolution and uranium
extraction from solid matrices." Separation and Purification Technology 257 (2021): 117950.
279
pubs.acs.org/IC Article
$
&
(
&
ACCESS Metrics & More Article Recommendations *
sı Supporting Information
+
ABSTRACT: Carbon and phosphorous are two primary elements common to the
"
'
(
bio-geosphere and are omnipresent in both biotic and abiotic arenas. Phosphonate
*
and carboxylate are considered as building blocks of glyphosate and humic
substances and constituents of the cellular wall of bacteria and are the driving
functionalities for most of the chemical interactions involving these two elements.
'
.
are ideal models to dig deep into for understanding the chemical interactions of the
two functional groups with metal ions. Phosphorous and carbon majorly exist as
Aquatic contamination is a major concern for uranium, and the presence of
'
complexing agents would alter the uranium concentrations in aquifers.
(
standing by density functional theory calculations was carried out to interpret the uranyl (UO22+) interaction with three
environmentally relevant phosphonocarboxylates, namely, phosphono-formic acid (PFA), phosphono-acetic acid (PAA), and
phosphono-propanoic acid (PPA). UO22+ forms 1:1 complexes with the three phosphonocarboxylates in the monoprotonated form,
-
having nearly the same stability, and the complexes [UO2(PFAH)], [UO2(PAAH)], and [UO2(PPAH)] involve chelate formation of
*
five, six, and seven membered rings, respectively, through the participation of an oxygen each from the carboxylate and phosphonate,
strengthened by an intra-molecular hydrogen bonding through the proton of the phosphonate moiety with uranyl oxygen. The
complex formations are favored both enthalpically and entropically, with the latter being more contributive to the overall free energy
of formation. The redox speciation showed an aqueous soluble complex formation over a wide pH range of 1−8. Electrospray
'
ionization mass spectrometry and extended X-ray absorption fine structure established the coordination modes, which are further
corroborated by density functional calculations. The knowledge gained from the present studies provide potential inputs in framing
(
&
(
*
&
the cleanup, sequestering, microbial, and bio-remediation strategies for uranyl from aquatic environments.
(
$
"
kinetics and reduction mechanism with in situ
Cite this: New J. Chem., 2017,
41, 15094 spectroelectrochemical investigation†
Sumit Kumara and Bhupendra S. Tomar ‡*a
Electrochemical methodology, viz. cyclic and square wave voltammetry, was used to investigate the
dominant species of UO22+–citrate at particular physiochemical conditions on the basis of their redox
behavior. The UO22+ in citrate media exists as UO22+, monomer [UO2Cit]ÿ, dimer [(UO2)2Cit2]2ÿ and polymeric
species. [UO2Cit]ÿ and [(UO2)2Cit2]2ÿ were found to be reduced in two irreversible one-electron steps,
with a chemical reaction coupled between them (ICI reduction mechanism), to a stable [UIVCit]+ species
at pH r 5 and via a single reduction step with two-electron transfers at pH 4 5. Chronopotentiometry
and chronoamperometry techniques were applied to corroborate the finding of an ICI reduction
mechanism, and to explore the kinetics of reduction by evaluating heterogeneous electron-transfer
kinetic (k0f,h and a) parameters. Using the Deford–Hume formula, the stability constant (log b) of mono-
meric species was calculated to be 6.8 0.11, which is in agreement with the literature. Spectroelectro-
Received 16th May 2017, chemical measurements were carried out to validate the chemical reaction interposed between the
Accepted 30th October 2017 two-electron transfer steps. Additionally, evidence of a new U(V) species, i.e., [UVO2Cit]2ÿ, in an aqueous
DOI: 10.1039/c7nj01701d system was obtained by in situ spectroelectrochemical measurements during electrolysis, and its stability
constant was obtained by cyclic voltammetry. ESI-MS studies on solutions of varying UO22+ : citrate
rsc.li/njc ratios and pH were used to confirm complex stoichiometries.
1. Introduction hence, there has been renewed interest in the redox chemistry
Speciation studies on uranium to predict its distribution and Uranium is a redox-sensitive element and its redox chem-
fate in the environment have been of considerable interest istry is exceptionally rich and diverse11 due to readily accessible
for quite some time.1–4 This becomes important because oxidation states ranging from III to VI and the sensitivity of the
anthropogenic activities such as milling and mining may lead redox potential to the coordination environment around the
to higher concentrations of uranium in the geosphere2,5 than metal ion.11,12 Investigation of the redox properties of uranium
that naturally abundant in soil and water (2–4 ppm and 3 ppb, in the presence of chelating agents13 is important from the point
respectively). Its radiological and toxicological effects have of view of its effect on uranium speciation involving solubility,
been the subject of interest to environmentalists.5–8 Hence, complexation, sorption, precipitation, bioavailability, and toxicity,14
knowledge of the speciation of uranium and its complexes is which are remarkably dependent on the oxidation states and
interesting with regard to understanding its migration behavior physiochemical conditions, viz. the redox potential (Eh) and pH.2
in an aquifer.9,10 Uranium mobility and transportation in the Citric acid (C3H5O(COOH)3) is a naturally occurring a-hydroxy
environment is strongly dependent on its oxidation states and, polycarboxylic acid that forms strong water-soluble complexes
with both U(VI) and U(IV) over a wide range of pH.15 It is a
a
Radioanalytical Chemistry Division, Bhabha Atomic Research Centre,
versatile chelating agent, and is a common constituent of
Mumbai-400085, India. E-mail: [email protected] domestic, industrial as well as nuclear wastes. With appropriate
b
Analytical Chemistry Division, Bhabha Atomic Research Centre, Mumbai-400085, steric arrangement it forms small chelate rings in the equatorial
India plane of the uranyl ion to give highly stable uranyl chelates.2
c
Fuel Chemistry Division, Bhabha Atomic Research Centre, Mumbai-400085, India
It exhibits relatively consistent removal efficiency, which has
† Electronic supplementary information (ESI) available. See DOI: 10.1039/
c7nj01701d
been used extensively for decontaminating contaminated sites
‡ Homi Bhabha National Institute, Bhabha Atomic Research Centre, Mumbai- in nuclear facilities. Uranyl ion complexation with citrate in
400085, India. environmental systems provides higher solubility, increased
15094 | New J. Chem., 2017, 41, 15094--15104 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
,QRUJDQLFD&KLPLFD$FWD ²
Research paper
a
Radioanalytical chemistry Division, Bhabha Atomic Research Centre, Mumbai 400085, India
b
Homi Bhabha National Institute, Anushaktinagar, Mumbai 400094, India
A R T I C LE I N FO A B S T R A C T
Keywords: Neptunium (Np) is a key element of concern for the safety aspects regarding the reprocessing and handling of
Neptunium spent nuclear fuel owing to its long half life (t1/2 = 2.14 × 106 years). Np can exist in several oxidation states in
Redox speciation aquatic environments depending on the surrounding conditions in which it exists. The +5 oxidation state of Np
Phenylphosphonic acid as NpO2+ is the most stable and least binding moiety that can migrate in aquatic arena to far away positions
Spectrophotometry
from its source of origin. Humic acids, which are omnipresent in aquifers, are the major source of binding for the
Density Functional Theory
actinides including Np to facilitate their migration and transportation phenomenon. Probing the systematic
knowledge of neptunium (Np), such as, speciation and coordination as function of its oxidation states in presence
of chelating agents is essential aspect of its solution chemistry. The present study aimed at investigating the
redox and complexation of Np with phenylphosphonic acid (PhPO3H2) (PPA) as mimic for binding of aromatic
phosphorus functionalities in humic substances by using electrochemical and UV–Visible absorption spectro-
scopy and density functional calculations. The cyclic voltammetric measurement indicates the complexation of
both oxidation states (V & VI) of Np with [PhPO3]2−. The presence and stability region (via Eh-pH plot) of new
species of Np-PhPO3 in two different oxidation states (V & VI) in aqueous solution at varying physicochemical
conditions are identified. The kinetics for reduction of Np complexes in different oxidation states are explored by
evaluation of heterogeneous electron-transfer kinetic (D0, ko and α) parameters by cyclic voltammetric results.
The spectral parameters, namely λmax and its molal extinction coefficient (ε) are obtained from UV–Visible
spectrophotometric measurements and are found to be 988.2 nm and 302 ± 8 mol−1 L cm−1 respectively. The
electrochemical as well as UV–Visible spectrophotometric measurements showed the formation of 1:1 complex
only and the stability constant (logβ) obtained for the same from both the methods found to be in agreement
with each other. Density Functional Theory (DFT) calculations are carried out to optimize the geometries of Np-
PhPO3 complex for the Np in +5 and +6 oxidation states with [PhPO3]2− in mono and bidentate mode to
identify the most probable coordination mode for complex formation.
⁎
Corresponding author at: Radioanalytical chemistry Division, Bhabha Atomic Research Centre, Mumbai 400085, India.
E-mail address: [email protected] (N. Rawat).
https://doi.org/10.1016/j.ica.2018.06.001
Received 5 March 2018; Received in revised form 31 May 2018; Accepted 1 June 2018
$YDLODEOHRQOLQH-XQH
(OVHYLHU%9$OOULJKWVUHVHUYHG
pubs.acs.org/journal/ascecg Research Article
%
Cite This: ACS Sustainable Chem. Eng. 2021, 9, 7846−7862 Read Online
#
%
&
ACCESS Metrics & More Article Recommendations *
sı Supporting Information
"
!
'
ABSTRACT: Research, based on the development of an economical innovative
&
+
.
-
its potential efficacy in diverse facets of the NFC, specifically as an elegant metal
hijacker for processing of uranium (U) matrices, designer solvent for green
&
metal, U3Si2, and (U, Pu)O2) were dissolved in Maline without any external
additives and cyclic voltammetry was performed to investigate the redox
speciation, viz., redox thermodynamics (Ep and Ef) and kinetic (D0, k0, and αn)
,
an insight into the molecular speciation, the structural analysis on UO3 dissolved
Molecular dynamics and density functional theory simulations were carried out to acquire diffusion, optimized structure, binding
energy, and molecular orbital diagram of U species in Maline, to corroborate the experimental results and to shed light on the
hydrogen-bond network in Maline with aqueous dilution. The interaction of uranyl with Maline was probed by luminescence,
&
absorption spectroscopy, and calorimetry titration. Green analysis methodology was developed based on Maline digestion followed
by voltammetric determination of U in nuclear material samples. Green chemistry metrics were evaluated to authenticate the greener
%
aspects of the present methodology. The present developed methodology of direct sequestration and analysis of U matrices
%
(
'
represents an appropriate replacement of the existing method, viz., hazardous acidic processing of U matrices followed by
'
biamperometry analysis.
KEYWORDS: deep eutectic solvent, Maline, direct sequestering, redox and molecular speciation, green analysis methodology
&
■ INTRODUCTION the nuclear fuel cycle (NFC), viz., various stages of processing
Nuclear energy, a mature technology and low carbon energy and reprocessing of metal and metal oxides, namely,
source, is going to be one of the major sources of energy due to dissolution, solvent extraction, recovery, and analysis. In the
increasing power demands and the possibility of reducing NFC, the current approaches of reprocessing fall into two
greenhouse gas concentration in the atmosphere.1−6 Uranium major categories as PUREX and pyrochemical processes.10,11
(U) is primarily utilized as a fuel source for nuclear power Both these processes have certain complexities and result in
generation and its increasing demand as fuel material in the technical challenges, viz., poor atom and energy efficiencies,
nuclear technology activity evokes processing of metal oxides
present in naturally occurring ores as well as in spent nuclear
fuel (SNF) for complete utilization of its energy potential.4−6 Received: March 4, 2021
However, the prime concern of nuclear energy is the economic Revised: May 12, 2021
and environmental issues associated with managing the SNF Published: May 27, 2021
wastes due to their long-term radiotoxicity.4,7,8 Therefore,
reprocessing and management of SNF are crucial to the future
of nuclear energy.4−6,9 Solvents occupy a strategic place within
A R T I C L E I N F O A B S T R A C T
Keywords: Two new deep eutectic solvents, different from hitherto known and based on common choline chloride, were
Deep eutectic solvents prepared. 1-methyl-3-decylimidazolium bromide mixed with hydrogen bond donors, malonic acid and digly-
Uranium oxides colic acid to form DES. Dissolution of uranium oxides (UO3 and UO2) was explored in both of them and char-
Redox speciation acterization of the resultant solutions was carried out by IR, TGA and UV–Vis spectroscopy. Redox speciation of
Cyclic voltammetry
dissolved uranium oxides was probed in these redox-active DES through cyclic voltammetry, differential pulse
In situ Spectroelectrochemistry
Theoretical simulations
voltammetry and in situ spectroelectrochemical measurements. Interestingly, the formation of uranium oxo
species was observed through the dissolution of UO3 in both the DES. The electrochemical characteristics
viz. redox thermodynamics (peak potential and formal redox potential), transport property (diffusion coeffi-
cient D0), heterogeneous electron transfer kinetic parameters (αn and k0) and mechanistic electron transfer
of the dissolved uranium species in two DES were investigated. The speciation of uranium was decoded
through the electronic absorption spectra of uranium in its lower oxidation states [U(V) and U(IV)] acquired
through in situ spectroelectrochemical electrolysis with varying cathodic potentials. Studies were also carried
out using Molecular Dynamics (MD) and density functional theory (DFT) simulations to authenticate the elec-
trochemical outcomes and to acquire the binding energy, optimized structure and molecular orbital diagram of
dissolved uranium species. Further, the MD simulation sheds light on the probable equatorial coordinating
atoms of dissolved uranium species. This is the first report, to the best of our knowledge, on imidazolium-based
DES and actinide ions.
⇑ Corresponding author at: Radiochemistry Division, Bhabha Atomic Research Centre, India.
E-mail address: [email protected] (M.S. Murali).
https://doi.org/10.1016/j.jelechem.2021.115752
Received 22 March 2021; Received in revised form 10 September 2021; Accepted 30 September 2021
Available online 05 October 2021
1572-6657/© 2021 Elsevier B.V. All rights reserved.
15th Biennial DAE BRNS Symposium Nuclear and Radiochemistry (NUCAR-2021), Feb 22-26, 2022, BARC, Mumbai
Probing speciation of Americium (Am) in deep eutectic solvent (DES) is indispensable for the
advancement of spent nuclear fuel (SNF) reprocessing based on liquid-liquid extraction as in the case
of SESAME and SANEX processes and for the growth of analytical tools including coulometry and
redox titration for Am analysis [1]. Am separation from SNF is highly challenging due to its similar
chemical and physical properties compared to the lanthanides (Ln). Selective extraction of Am vs. Ln
through judicious ligand design is the mainstream approach for Am separation but this approach
requires multiple steps and can cause more down-stream complication. Selective oxidation of Am(III)
to its higher oxidation state is another approach which could facilitate its separation from lanthanides.
Therefore, development of electrochemical methodology for selective oxidation of Am to its higher
oxidation which could facilitate its separation from curium (Cm) as well as lanthanides has become a
recent research interest. In this context, DESs, an interesting new class of green and eco-sustainable
solvents which shown their usefulness as environmentally benign alternatives to conventional organic
solvents and ionic liquid, would be a suitable choice. DESs are a kind of eutectic mixture of hydrogen
bond acceptors (HBAs) and hydrogen bond donors (HBDs) with a considerably lower melting point
than its original components. DESs have some fascinating and tailored properties (large solubility with
metal oxides, wide potential windows, and the possibility of tuning the properties of DES by changing
the functional groups of solvent components) which make them promising electrolytes for
electrochemical processes and a good candidate for selective coordination of actinides. Owing to
unique physicochemical properties of DES, the speciation of actinides in DES might be different
compare to that in aqueous/ionic liquid medium. Literature survey indicates the use of DES as co-
solvent for Am separation [2] however, no studies have been reported on the speciation of Am in DES.
The aim of the present study is to describe the speciation of Am in two DES containing mainly oxygen
atoms as functional moiety viz., Choline dihydrogen Citrate-Malonic acid (CDHC-MA) and Choline
Chloride-Malonic acid (CC-MA). The Am-oxide was dissolved in both DES and the dissolved
solutions were characterized by UV-visible spectroscopy [Table 1]. The results indicate Am (III)-DES
complex species formation in both DES through dissolution process. In depth analysis on the
electrochemical oxidation of Am(III) in both DES were carried out by cyclic and differential pulse
voltammetry, indicates (i) one oxidative peak with one electron transfer as Am(III) to Am(IV) and
hence Am(III) can be easily oxidized to stable Am(IV), (ii) Am(IV) is a stable oxidation state in both
DES and (iii) Am(III)-Am(IV) conversion has much lower oxidation potential (~ 0.68 V) in DES
compared to that in aqueous medium (i.e., 2.6 V vs. SCE). The redox parameters (peak potential and
diffusion coefficient) of above oxidation process have been reported in Table 1. Theoretical
calculations were performed to corroborate the electrochemical results and to obtain the optimized
structure and energetic [Table 1] of Am in both DES [1:2 type
complexes formed].
Table 1. Experimentally and theoretically obtained parameters of Am in DES
References: 1. S. Picart et al., ATALANTE-2004 P2-P4. Nimes, France, June 21-25 (2004).
2. G. Colombo Dugoni et al., ACS omega, 6(5), 3602 (2021).