Quick viewing(Text Mode)

Lanthanide-Polyaminopolycarboxylate Complexation Kinetics in High Lactate Media

Lanthanide-Polyaminopolycarboxylate Complexation Kinetics in High Lactate Media

LANTHANIDE-POLYAMINOPOLYCARBOXYLATE COMPLEXATION KINETICS IN

HIGH LACTATE MEDIA: INVESTIGATING THE AQUEOUS PHASE OF TALSPEAK

By

DEREK MACKENZIE BRIGHAM

A dissertation submitted in partial fulfillment of the requirements for the degree of

DOCTOR OF PHILOSOPHY

WASHINGTON STATE UNIVERSITY Department of Chemistry

May 2013

© Copyright by DEREK MACKENZIE BRIGHAM, 2013 All Rights Reserved

© Copyright by DEREK MACKENZIE BRIGHAM, 2013 All Rights Reserved

To the Faculty of Washington State University:

The members of the Committee appointed to examine the dissertation of DEREK MACKENZIE BRIGHAM find it satisfactory and recommend that it be accepted.

______Kenneth L. Nash, Ph.D., Chair

______Scot E. Wherland, Ph.D.

______Sue B. Clark, Ph.D.

______Jeremy J. Lessmann, Ph.D.

ii

ACKNOWLEDGMENTS

“Faith in your partner, your fellow men, your friends, is very important, because without it there's no mutual component to your relationship, and relationships are important. So faith plays an important role, but faith in people you don't know, faith in religious or political leaders or even people on stages, people who are popular in the public eye, you shouldn't have faith in those people. You should listen to what they have to say and use it. It might give you some ideas on how to view the world, but ultimately you have to base your views on evidence. Evidence comes from your own eyes and ears.” Dr. Greg Graffin

When I first arrived at WSU to begin my undergraduate studies I had very different ambitions and could not have predicted where I would end up. I certainly would not have guessed that I would still be in Pullman nearly ten years later. While this journey has been long and at times arduous, I am glad for all the experiences it has afforded me.

I want to thank my friends and family back home in Oregon for always being there for me. I certainly would not be where I am today without the ever present love and support of my parents, Larry and Marie Brigham. I am especially grateful of their patience with me throughout my life. I am deeply appreciative for all they have done for me, and hope to always make them proud. When I would come home to visit my friends made it feel like I had never left. This always gave me something to look forward to, and left me refreshed and ready to return. My life would not be the same without them.

My time at WSU has left me with some of the closest friends I could have ever asked for.

The ones I have trained with, studied with, laughed, cried, suffered, and been triumphant with, I will never forget them. It is rare to make this kind of connection, and it is something I will always cherish.

iii

I wish to express my deepest gratitude to my committee members. I would first like to thank Dr. Sue Clark, not only for being on my committee but for helping to set me on the path that has brought me to this point. If not for her involvement in the undergraduate chemistry program I would not have been introduced to radiochemistry. Dr. Jeremy Lessmann has been a welcome presence throughout both my undergraduate and graduate experience. It has been a pleasure to work for and to work with him. I am especially grateful of Dr. Scot Wherland. Not only has he been an excellent teacher, who taught me most of what I know about kinetics, but he was always welcoming and insightful when I would come to his door unannounced to discuss my results. I would not have come this far without his assistance. And finally I wish to thank Dr.

Ken Nash, who has been an excellent mentor, teacher, and friend. I truly believe I would not have made it through the graduate program without his belief in, and patience with, me.

I wish to make special mention of the late James “Sully” Sullivan. He was the one to first foster my interest in chemical kinetics, and he showed me what a true passion for science was.

As the last student Sully took on, I hope to honor his memory and pass on his passion for research and discovery.

iv

LANTHANIDE-POLYAMINOPOLYCARBOXYLATE COMPLEXATION KINETICS IN

HIGH LACTATE MEDIA: INVESTIGATING THE AQUEOUS PHASE OF TALSPEAK

Abstract

by Derek Mackenzie Brigham, Ph.D. Washington State University May 2013

Chair: Kenneth L. Nash

In advanced nuclear fuel reprocessing schemes, the TALSPEAK (Trivalent Actinide

Lanthanide Separation with Phosphorus-reagent Extraction from Aqueous Komplexes) process has been proposed as a means to separate Am and Cm from the . One significant limitation of the TALSPEAK process is slow phase transfer kinetics of the lanthanides to the organic phase. Increasing the lactic acid buffer concentration is found to improve the solvent extraction kinetics. However, concentrations of greater than 1 M are necessary to achieve rates of mass transfer fast enough for TALSPEAK to be applicable on an industrial scale. The

TALSPEAK process employs diethylenetriaminepentaacetic acid (DTPA) as an aqueous phase complexant to selectively bind to the actinides and prevent their extraction, however, DTPA also binds with the lanthanides. Understanding the mechanism of the interaction between DTPA and the lanthanides in high total lactate will help to explain the accelerative effect of increased total lactate on TALSPEAK mass transfer rates.

This dissertation describes the homogeneous aqueous complexation kinetics of the lanthanides Pr3+, Nd3+, Sm3+-Lu3+ and the polyaminopolycarboxylate DTPA,

v ethylenediaminetetraacetic acid (EDTA), and (hydroxyethyl)ethylenediaminetriacetic acid

(HEDTA) in 1 M total lactate aqueous media similar to that found in the aqueous phase of a

TALSPEAK separation system. Temperature studies on the interactions of select lanthanides with DTPA, EDTA, and HEDTA were performed to obtain activation parameters associated with the complex formation reaction. Additional studies on the interaction of Eu3+ with DTPA were performed under varying total lactate conditions at several different constant concentrations of lactate ion and pH values to determine the mechanistic role of the species in the lactate/lactic acid buffer system. Kinetic data were obtained using the method of equilibrium perturbation by displacement via stopped-flow spectrophotometry employing the colorimetric dye arsenazo III.

This work provides an increased understanding of lanthanide interactions with polyaminopolycarboxylate ligands in TALSPEAK-like aqueous media. From the insights gained in these studies, a possible explanation for the accelerative effect of lactate on TALSPEAK phase transfer rates is proposed. The overall conclusion of this work is that under high concentrations of total lactate the lactate ion governs the aqueous phase complexation kinetics.

vi

TABLE OF CONTENTS

ACKNOWLEDGMENTS ...... iii

ABSTRACT ...... v

LIST OF TABLES ...... ix

LIST OF FIGURES ...... xi

CHAPTER

1. INTRODUCTION ...... 1

Electricity in the U.S ...... 1

Nuclear Power ...... 1

Nuclear Fuel Reprocessing ...... 2

TALSPEAK ...... 4

Kinetics of Extraction in TALSPEAK ...... 6

Research Aims ...... 9

2. EXPERIMENTAL ...... 13

Reagents ...... 19

Procedure ...... 20

Trans-lanthanide Study ...... 21

Temperature Studies ...... 22

Total Lactate Experiments ...... 22

AAIII Independence...... 23

3. EXPERIMENTAL RESULTS...... 28

DTPA Lanthanide Series Study ...... 29

Temperature Study ...... 33

vii

EDTA Lanthanide Series Study ...... 35

Temperature Study ...... 39

HEDTA Lanthanide Series Study ...... 41

Temperature Study ...... 45

High Ligand Concentration Study ...... 47

4. TRANS-LANTHANIDE STUDY DISCUSSION ...... 53

Trans-lanthanide Trends ...... 53

Isokinetic Relationship...... 59

Calculated Stability Constants ...... 63

Lactate Complexation ...... 66

High Ligand Concentration...... 75

5. RESOLVING THE MECHANISTIC ROLE OF THE LACTATE BUFFER ...... 78

Constant pH ...... 80

Constant Lactate Ion ...... 84

Role of Lactic Acid ...... 88

High [DTPA] Under Constant Lactate Ion Conditions ...... 90

Data Fitting ...... 94

Conclusion ...... 100

CONCLUSION ...... 102

APPENDIX A ...... 112

viii

LIST OF TABLES Chapter 3 3.1 Lanthanide-DTPA second order rate constant of complex formation. 1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C ...... 30

3.2 Lanthanide-DTPA first order rate constant of complex dissociation. 1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C ...... 30

3.3 Activation parameters for the Ln-DTPA complexation reaction in 1.0 M NaLac 0.9 M HClO4 and 2 M ionic strength. Additionally the second order rate constant of complex formation at 25 °C is listed ...... 35

3.4 Lanthanide-EDTA second order rate constant of complex formation. 1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C ...... 37

3.5 Lanthanide-EDTA first order rate constant of complex dissociation. 1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C ...... 37

3.6 Activation parameters for the Ln-EDTA complexation reaction in 1.0 M NaLac 0.9 M HClO4 and 2 M ionic strength. Additionally the second order rate constant of complex formation at 25 °C is listed ...... 41

3.7 Lanthanide-HEDTA second order rate constant of complex formation. 1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C ...... 43

3.8 Lanthanide-HEDTA first order rate constant of complex dissociation. 1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C ...... 43

3.9 Activation parameters for the Ln-HEDTA complexation reaction in 1.0 M NaLac 0.9 M HClO4 and 2 M ionic strength. Additionally the second order rate constant of complex formation at 25 °C is listed ...... 47

Chapter 4 4.1 Activation enthalpies, entropies, and rate constants at 25 °C for the complex formation reaction between the Ln’s and PAPC ligands in 1.0 M NaLac, 0.9 M HClO4 and 2 M ionic strength ...... 60

4.2 Calculated kinetic equilibrium constants from the rate constants obtained in1.0 M NaLac 0.9 M HClO4 2 M ionic strength. Thermodynamic calculated equilibrium constants calculated from literature data at 0.1 M ionic strength and 25 °C ...... 68

4.3 Average number of lactate ions coordinated to Ln3+ in 1.0 M NaLac at a given pH ...... 69

ix

Chapter 5

5.1 Rate constants of complex formation (kf) and dissociation (kd) under constant pH with different total lactate conditions ...... 83

5.2 Rate constants of complex formation (kf) and dissociation (kd) under constant lactate ion concentrations with different total lactate conditions. Respective pH - - values at 1.0, 0.9, and 0.8 M Lactot; 0.1 M Lac : 2.72, 2.77,and 2.83; 0.2 M Lac : 3.07, 3.13, and 3.19; 0.3 M Lac-: 3.30, 3.37, and 3.45 ...... 87

Appendix A.1 Upper and lower sensitivity limits of calculated fitting parameters resulting in a 20% increase in the residual sum of squares. *Zero value for this parameter results in a 13% increase in RSS...... 114

x

LIST OF FIGURES Chapter 1 1.1 Structure of di(2-ethylhexyl) phosphoric acid (HDEHP) ...... 5

1.2 Structure of lactic acid ...... 5

1.3 Structure of diethylenetriaminepentaacetic acid (DTPA) ...... 6

1.4 Lanthanide extraction at 30 minutes by 0.3 M HDEHP in n-dodecane from 0.05 M DTPA and () 0, (●) 0.1, and (■) 1.0 M total lactate at pH 3.5. Reproduced from [8] ...... 7

1.5 Rate of complexation for the reaction of La3+ with DTPA in 0.1 (■), 0.2 (●), and 0.3 () M total lactate at pH 3.5. Reproduced from [9] ...... 8

1.6 Structure of ethylenediaminetetraacetic acid (EDTA) ...... 9

1.7 Structure of (hydroxyethyl)ethylenediaminetriacetic acid (HEDTA) ...... 10

Chapter 2 2.1 Structure of the Arsenazo III...... 15

2.2 Change in the visible spectrum upon mixing of EuAAIII with DTPA. 0.05 mM Eu 0.0125 mM AAIII and 8 mM DTPA in 1.0 M NaLac, pH 2.6, 2 M ionic strength 25 °C ...... 16

2.3 Complexation of 0.05 mM Gd and EDTA 1.0 M NaLac 0.9M HClO4 2 M ionic strength 25 °C with [AAIII] (■) 0.0125 mM and (●) 0.025 mM at pH 2.7 and 2.8 respectively ...... 24

Chapter 3 3.1 Observed rate constants of the reaction between Eu3+ and DTPA in 1.0 M NaLac 0.9 (■), 0.85 (●), and 0.8 () M HClO4 2 M ionic strength at 25 °C ...... 29

3.2 Ln-DTPA kf values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M HClO4, 2 M ionic strength at 25 °C ...... 31

3.3 Ln-DTPA kd values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M HClO4, 2 M ionic strength at 25 °C ...... 32

3.4 Temperature dependence of the reaction between Eu3+ and DTPA 1 M NaLac 0.9 M HClO4 2 M ionic strength at 25 °C (■), 20 °C (●), 15 °C () and 10 °C ()...... 34

3.5 Observed rate constants of the reaction between Eu3+ and EDTA in 1.0 M NaLac and 0.9 (■), 0.85 (●), and 0.8 () M HClO4 2 M ionic strength at 25 °C ...... 36

xi

3.6 Ln-EDTA kf values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M HClO4, 2 M ionic strength at 25 °C ...... 38

3.7 Ln-EDTA kd values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M HClO4, 2 M ionic strength at 25 °C ...... 39

3.8 Temperature dependence of the reaction between Eu3+ and EDTA 1 M NaLac 0.9 M HClO4 2 M ionic strength at 30 °C (■), 25 °C (●), 20 °C () and 15 °C ()...... 40

3.9 Observed rate constants of the reaction between Eu3+ and HEDTA in 1.0 M NaLac and 0.9 (■), 0.85 (●), and 0.8 () M HClO4 2 M ionic strength at 25 °C ...... 42

3.10 Ln-HEDTA kf values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M HClO4, 2 M ionic strength at 25 °C ...... 44

3.11 Ln-HEDTA kd values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M HClO4, 2 M ionic strength at 25 °C ...... 45

3.12 Temperature dependence of the reaction between Eu3+ and HEDTA 1 M NaLac 0.9 M HClO4 2 M ionic strength at 33 °C (■), 33 °C (●), 27 °C () and 25 °C ()...... 46

3+ 3.13 Complexation of Eu with (■) DTPA, (●) EDTA, and () HEDTA in 1.0 M NaLac, 0.9 M HClO4, and 2 M ionic strength at 25 °C. Lines added as a visual guide ...... 48

3.14 Observed rate constants of the reaction between Eu3+ and DTPA in 1.0 M NaLac 0.9 (■), 0.85 (●), and 0.8 () M HClO4 2 M ionic strength at 25 °C. Lines added as a visual guide...... 49

Chapter 4 4.1 Rate of complex formation across the lanthanides with (■) DTPA, (●) EDTA, and () HEDTA in 1.0 M NaLac 0.9 M HClO4 2 M ionic strength at 25 °C ...... 54

4.2 Rate of complex formation of the lanthanides with (■) DTPA, and (●) EDTA in 0.3 M Lactot pH 3.6 (reproduced from [1]); (□) DTPA and (○) EDTA in 1.0 M Lactot pH 3.0 at 25 °C ...... 56

4.3 Rate of complex dissociation across the lanthanides with (■) DTPA, (●) EDTA, and () HEDTA in 1.0 M NaLac 0.9 M HClO4 2 M ionic strength at 25 °C ...... 57

4.4 Rate of complex formation of the lanthanides with (■) DTPA, and (●) EDTA in 0.3 M Lactot pH 3.6; (□) DTPA and (○) EDTA in 1.0 M Lactot pH 3.0 at 25 °C ...... 58

4.5 Isokinetic relationship for the complexation reaction between the lanthanides Nd3+, 3+ 3+ 3+ 3+ 3+ Eu , Tb , Ho , Tm , and Lu with (■) DTPA, (●) EDTA, and () HEDTA ...... 61

xii

4.6 Rate constant of complex formation of Eu3+ with (■) DTPA and (●) EDTA as a function of the average number of lactate ions coordinated to Eu3+ prior to mixing ...... 69

4.7 Rate constant of complex formation between the lanthanides with DTPA as a function of the average number of coordinated lactate ions. 1.0 M NaLac, pH 2.6, at 25 °C ...... 70

4.8 Rate constant of complex formation between the lanthanides with EDTA as a function of the average number of coordinated lactate ions. 1.0 M NaLac, pH 2.6, at 25 °C ...... 71

4.9 Rate constant of complex formation between the lanthanides with HEDTA as a function of the average number of coordinated lactate ions. 1.0 M NaLac, pH 2.6, at 25 °C ...... 72

4.10 Rate constant of complex formation with ligand vs. average number of bound lactates in (■) 0.3 M total lactate pH 3.6 with the lanthanides La3+-Lu3+ [1] and (●) 1.0 M total lactate pH 3.0 with the lanthanides Pr3+-Lu3+ ...... 73

4.11 Rate constant of complex dissociation between the lanthanides with EDTA as a function of the average number of coordinated lactate ions. 1.0 M NaLac, pH 2.6, at 25 °C ...... 74

Chapter 5 5.1 Speciation of DTPA in 2 M ionic strength ...... 79

5.2 Observed rate constants for the reaction between Eu3+ with DTPA under a constant pH of 2.72 in 1.0 (■), 0.9 (●), and 0.8 () M total lactate ...... 81

5.3 Observed rate constants for the reaction between Eu3+ with DTPA under a constant pH of 3.07 in 1.0 (■), 0.9 (●), and 0.8 () M total lactate ...... 82

5.4 Observed rate constants for the reaction between Eu3+ with DTPA under a constant pH of 3.30 in 1.0 (■), 0.9 (●), and 0.8 () M total lactate ...... 83

5.5 Observed rate constants for the reaction between Eu3+ with DTPA under a - constant [Lac ] of 0.1 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 2.72, 2.77, and 2.83 respectively ...... 85

5.6 Observed rate constants for the reaction between Eu3+ with DTPA under a - constant [Lac ] of 0.2 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 3.07, 3.13, and 3.19 respectively ...... 86

5.7 Observed rate constants for the reaction between Eu3+ with DTPA under a - constant [Lac ] of 0.3 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 3.30, 3.37, and 3.45 respectively ...... 87

xiii

5.8 Observed rate constants for the reaction between Eu3+ with DTPA under the conditions of constant pH of 3.07 (left) and a constant [Lac-] of 0.2 M (right) in 1.0 (■), 0.9 (●), and 0.8 () M total lactate ...... 89

5.9 Observed rate constants for the reaction between Eu3+ with DTPA under a - constant [Lac ] of 0.1 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 2.72, 2.77, and 2.83 respectively ...... 91

5.10 Observed rate constants for the reaction between Eu3+ with DTPA under a - constant [Lac ] of 0.2 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 3.07, 3.13, and 3.19 respectively ...... 92

5.11 Observed rate constants for the reaction between Eu3+ with DTPA under a - constant [Lac ] of 0.3 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 3.30, 3.37, and 3.45 respectively ...... 93

5.12 Experimental data and calculated fits for the constant [Lac-] condition of 0.1 M 2 in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with respective R values for the fits of 0.976, 0.999, and 0.978. Fitting parameters – a: 3813; b: 2.29•10-3; c: 1.22 ...... 96

5.13 Experimental data and calculated fits for the constant [Lac-] condition of 0.2 M 2 in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with respective R values for the fits of 0.964, 0.994, and 0.938. Fitting parameters – a: 4173; b: 4.57•10-3; c: 2.19 ...... 97

5.14 Experimental data and calculated fits for the constant [Lac-] condition of 0.3 M 2 in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with respective R values for the fits of 0.968, 0.998, and 0.967. Fitting parameters – a: 4961; b: 4.32•10-3; c: 4.62 ...... 98

xiv

DEDICATION

In loving memory of Martha Ann Brigham

xv

Chapter 1

Introduction

Electricity in the U.S.

In 2010 the United States accounted for 21% of global electricity consumption. In 2011

68% of the electricity generated in the U.S. was through combustion of fossil fuels.1 With the rising concern over the adverse effects of carbon dioxide emission to the atmosphere on the global climate, there is increased pressure for implementing alternative methods of power production. With federal government incentives, wind power generation has seen increased use in the past several years, accounting for 3% of total power generation in 2011, up from 0.8% in

2007.1 While renewable energy sources such as wind and solar will play an important role in the reduction of reliance on fossil fuels, the intermittent nature of these renewable energy sources means their use on a large energy grid will require complementary alternative forms of power generation, which often is supplied by a conventional thermal power plant run on coal, oil, or natural gas.

Nuclear Power

Of presently available large capacity options, the best means available to reduce CO2 emissions from electricity generation is by increased implementation of nuclear power. In 2011 nuclear power accounted for 19% of the electricity produced in the United States.1 It is the only

CO2 emission-free method of production that is both well established and capable of replacing a significant percentage of the power generated by fossil fuel combustion. Conventional nuclear power reactors moderated and cooled by light water generate electricity by utilizing uranium fuel

1 that has been enriched to between three and five percent 235U. The neutrons generated in the fissioning of 235U (and later by 239Pu that is bread from 238U) maintain the chain reaction. This exothermic process hears water to produce steam that runs turbines that generate electricity.

Besides reducing the amount of carbon dioxide released to the atmosphere as a result of power production, nuclear power generation could reduce the amount of radioactive material released to the environment from electricity production. In 2011 42% of the electricity generated in the U.S. was done by burning coal.1 Coal contains small quantities of uranium, thorium, and their radioactive daughter products. In a study by McBride et. al. it was found that populations received a greater dose of radiation from coal-fired power plants than from nuclear power plants.2

Nuclear Fuel Reprocessing

The main drawback to nuclear power generation is the highly radioactive waste that remains after the uranium fuel elements have gone through one (typically four year) cycle in the reactor. The vast majority of this material (>95%) can be recycled for additional power generation, however the remaining material still presents a substantial challenge. Without treatment the long term radiotoxicity hazard requires sequestration of the used fuel on a time scale reaching hundreds of thousands of years.3 However, this time span can be greatly reduced by reprocessing the spent nuclear fuel.

Although no reprocessing of used fuel is performed in the U.S., the U.K., Russia, Japan, and France recycle their spent nuclear fuel, mainly by some form of the PUREX process. The

PUREX (Plutonium Uranium Recovery by EXtraction) process was developed by the U.S. during the Manhattan Project, primarily as a means for the production of plutonium for use in

2 nuclear weapons. The PUREX liquid-liquid separation scheme employs the extractant tributyl phosphate (TBP) to separate uranium and plutonium from the fission products and other actinides.3 Solutions of 30% TBP in kerosene are contacted with 3-4 M nitric acid aqueous solutions of the dissolved spent nuclear fuel. In the first cycle more than 99.8% of uranium and plutonium are extracted (in hexavalent and tetravalent oxidation states respectively) leaving more than 99% of the fission products in the aqueous raffinate. In later stages, plutonium is reduced to the trivalent state by contact with a solution containing a suitable reductant, such as

U(IV). The Pu(III) then strips to the aqueous phase, separating from the uranium. At this point, the PUREX solvent extraction method produces a stream of clean weapons grade plutonium and as such raises concerns over nuclear proliferation.

UREX+, an advanced nuclear reprocessing scheme developed at Argonne National

Laboratory, employs a suite of solvent extraction methods to reduce the risk of proliferation as well as the length of time the waste is hazardous to the environment.4 The first step is the UREX

(Uranium Extraction) process, a modification of the PUREX process. UREX avoids producing pure plutonium by adding the reductant acetohydroxamic acid to the aqueous feed. The Pu(IV) is reduced to Pu(III) which is poorly extracted and remains in the aqueous phase while the uranium is extracted to the organic phase by TBP. The UREX aqueous raffinate is then introduced to the the CCD-PEG segment. A mixture of chlorinated cobalt dicarbollide (CCD) and polyethylene glycol (PEG) extract cesium and strontium respectively, which are the dominant fission products.

The CCD-PEG aqueous raffinate then moves on to the NPEX stage. The NPEX (Neptunium

Plutonium Extraction) process thermally destroys the acerohydroxamic acid reductant/complexant introduced in the UREX stage and increases the concentration of nitric acid to convert plutonium and neptunium to the extractable tetravalent oxidation state. Pu(IV)

3 and Np(IV) are then extracted by TBP in a process similar to that of UREX and PUREX. The aqueous raffinate of NPEX then moves on to the TRUEX (TRansUranium EXtraction) segment.

TRUEX employs the carbamoyl organophosphorus extractant octyl(phenyl)-N,N- diisobutylcarboylmethyl-phosphine oxide (CMPO) and TPB to extract americium, curium, and the fission product lanthanides from the remaining fission products.

The product stream of TRUEX contains the most important isotopes contributing to the long term radiotoxicity of the spent fuel, the actinides americium and curium. Removing these elements and returning them to reactors for transmutation to shorter lived isotopes significantly reduces the amount of time the fuel elements need to be isolated from the biosphere. One obstacle to this method, however, is the presence of the trivalent lanthanides, which act as neutron poisons and reduce the efficiency of the actinide transmutation. Unfortunately the trivalent lanthanides have chemical behavior very similar to the trivalent actinides and are quite difficult to separate from one another. The separation of these elements has been a major factor in nuclear fuel cycle research.

TALSPEAK

One method for separating the trivalent actinides from the trivalent lanthanides is the solvent extraction system known as the TALSPEAK (Trivalent Actinide Lanthanide Separation with Phosphorus-reagent Extraction from Aqueous Komplexes) process. Developed at Oak

Ridge National Laboratory in the 1960’s by Weaver and Kappelman5, the TALSPEAK process employs the liquid cation exchange extractant di(2-ethylhexyl) phosphoric acid (HDEHP) (figure

1.1) to selectively extract the trivalent lanthanides into the organic phase from aqueous solutions.

The pH of the aqueous medium in conventional TALSPEAK is maintained between pH 3.0 and

4

3.5 using a lactic acid buffer (figure 1.2). The trivalent actinides are held back in the aqueous phase by the complexing agent diethylenetriaminepentaacetic acid (DTPA) (figure 1.3). While the trivalent actinides have very similar chemistry to that of the trivalent lanthanides, the f- orbitals of the actinides extend slightly out of the d-orbital subshell while the lanthanide f- orbitals are completely shielded. This gives the actinides some soft Lewis-acid character which leads to greater interactions with soft Lewis-base electron donors, in this case the amine donor atoms of DTPA. The TALSPEAK process takes advantage of this characteristic of the actinides to achieve separation between the actinides and lanthanides by using DTPA as an aqueous hold- back reagent. The softer nitrogen-donors of DTPA interact with the trivalent actinides more strongly than the lanthanides. As metal-DTPA complexes are non-extractable, and DTPA will preferentially bind to the actinides, the lanthanides are able to be selectively extracted by

HDEHP to the organic phase.

Figure 1.1: Structure of di(2-ethylhexyl) phosphoric acid (HDEHP).

Figure 1.2: Structure of lactic acid.

5

Figure 1.3: Structure of diethylenetriaminepentaacetic acid (DTPA).

Kinetics of Extraction in TALSPEAK

The TALSPEAK process has been subject to extensive study since its introduction.6

Nevertheless, the chemistry underpinning the process is not yet completely understood.

Although TALSPEAK seems to be quite promising for application to nuclear fuel reprocessing, the process is not without limitations. One such complication is that the phase transfer kinetics of

TALSPEAK are very slow at low concentrations of the lactic acid buffer. It has been observed that increasing the concentration of total lactate in the system increases the rate of lanthanide extraction.7 The data in figure 1.4 show the effect of increasing lactate on the extraction kinetics in TALSPEAK across the lanthanide series.8 For the phase transfer kinetics to be fast enough for

TALSPEAK to be applicable as an industrial process, concentrations of lactate greater than 1.0

M are necessary. The need for such high concentrations of carboxylic acid buffer have been established as one contributing factor to the complexity of the TALSPEAK process.

6

105

104

103

2

10 D

101

100

10-1 0.84 0.86 0.88 0.90 0.92 0.94 0.96 0.98 1.00 1.02 1.04 Inverse ionic radius (Å-1) CN=8

Figure 1.4: Lanthanide extraction at 30 minutes by 0.3 M HDEHP in n-dodecane from 0.05 M DTPA and () 0, (●) 0.1, and (■) 1.0 M total lactate at pH 3.5. Reproduced from [8].

Increasing the concentration of lactic acid is seen to increase the rate of lanthanide extraction in the TALSPEAK process, however the reason for this accelerative effect is unclear.

An investigation of the complexation kinetics in TALSPEAK-like aqueous media, performed by

Nash et. al., could provide insight to the lactate effect on the extraction kinetics.9 However, the results of this study seem to present further complications. In a pseudo first-order study of the complexation kinetics between La3+ and DTPA in various concentrations of total lactate at pH

-1 -1 3.5, it was found that the rate constant of complex formation, kf (M s ), is constant at different concentrations of total lactate. Figure 1.5 reproduces these results, where under pseudo first- order conditions the slope gives the kf value and the intercept gives the kd value (derivation of this relationship provided in chapter 2). It was noted that the rate constant of complex

-1 dissociation, kd (s ), decreases with increasing total lactate.

7

0.06

0.05

0.04

) -1

0.03

(ms

obs k 0.02

0.01

0.00 0.0000 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 DTPA (M)

Figure 1.5: Rate of complexation for the reaction of La3+ with DTPA in 0.1 (■), 0.2 (●), and 0.3 () M total lactate at pH 3.5. Reproduced from [9].

The observation that the Ln-DTPA rate constant of complex dissociation decreases with an increase in total lactate concentration indicates a greater kinetic stability of the complex.

Greater kinetic stability of the DTPA complexes means the lanthanides will spend more time bound to the aqueous complexant, which is a non-extractable species and therefore should slow the rate of phase transfer in a TALSPEAK-like process. This observed effect of increased concentration of total lactate on the aqueous complexation kinetics appears to be in conflict with the observation that higher concentrations of total lactate increase the rate of lanthanide extraction in TALSPEAK. Clearly, more complex interactions must be at work.

8

Research Aims

The apparent conflict between the experimental results of increased total lactate concentration on the kinetics of lanthanide extraction and the aqueous complexation kinetics of the lanthanides motivated behind the work that is to follow. In examining the experiments on the extraction kinetics to the aqueous complexation kinetics, it is noteworthy that the aqueous lanthanide-DTPA complexation kinetics experiments occur in 0.1-0.3 M total lactate. This range is considerably lower than the 1.0 M concentration where the most significant effect on the extraction kinetics is observed. In order to be more representative of the TALSPEAK aqueous phase, the complexation kinetics experiments need to be performed in solutions of higher total lactate concentration. Therefore, the work that is to follow investigates the complexation kinetics of the lanthanides with polyaminopolycarboxylate (PAPC) ligands in 1.0 M total lactate by stopped-flow spectrophotometry. Three PAPC ligands selected for their decreasing denticity and bonding strength will be used in the following complexation kinetics studies, DTPA (figure 1.3), ethylenediaminetetraacetic acid (EDTA) (figure 1.6), and

(hydroxyethyl)ethylenediaminetriacetic acid (HEDTA) (figure 1.7). The systematic changes in the structures and binding pockets of these ligands should provide some insight to the effect of the different ligands on the complexation kinetics.

Figure 1.6: Structure of ethylenediaminetetraacetic acid (EDTA).

9

Figure 1.7: Structure of (hydroxyethyl)ethylenediaminetriacetic acid (HEDTA).

Temperature studies of the complexation kinetics of these PAPC ligands with the lanthanides will provide a great deal of information on how the mechanism of the complexation reaction changes between the different ligands. Completing these studies across the lanthanide series will allow for correlations to be developed between rates of complexation, ligand denticity, thermodynamic stability and the cation size in transiting the lanthanide series, as the radii shrink by about 20% across the. As the total lactate in these systems are comprised of the lactic acid and lactate ion species and their concentrations are dependent on the pH of the system, these complexation kinetics were investigated under three different acidity conditions to gain a cursory understanding of the effect of these different species of lactate on the kinetics.

However, to successfully interpret the kinetic role of the lactic acid, lactate ion, and hydrogen ion species in these systems, more complex experiments with varying concentrations of total lactate will be required. Understanding the roles of these different species allows for mechanistic interpretation of the aqueous complexation reactions that take place in TALSPEAK.

The goal of the experiments to follow is to gain a better understanding of the aqueous complexation kinetics in high lactate media so that the conflict between the results of previous complexation kinetics and extraction kinetics experiments may be resolved. Additionally, a greater mechanistic understanding of how the lanthanides interact with the aqueous hold-back

10 reagent DTPA in high lactate potentially provides insight as to how the complicated TALSPEAK extraction process operates.

11

References

1. U.S. Energy Information Administration. http://www.eia.gov/, February 2013.

2. McBride, J. P.; Moore, R. E.; Witherspoon, J. P.; Blanco, R. E. Radiological Impact of Airborne Effluents of Coal and Nuclear Plants. Science, 1978, 202, 1045.

3. Choppin, G.R.; Liljenzin, J-O.; Rydberg J. Chapter 21: Nuclear fuel cycle. Radiochemistry and Nuclear Chemistry, 3rd edition; Butterworth-Heinemann: Woburn, MA, 2002; 583-641.

4. Vandegrift, G. F.; Regalbuto, M. C.; Aase, S. C.; Bakel, A. J.; Battisti, T. J.; Bowers, D. L.; Byrnes, J. P.; Clark, M. A.; Emery, J. W.; Falkenberg, J. R.; Gelis, A. V.; Pereira, C.; Hafenrichter, L. D.; Tsai, Y.; Quigley, K. J.; Vander Pol, M. H. Designing and demonstration of the UREX+ process using spent nuclear fuel. Proceedings of Atatlante 2004: Advances of Future Nuclear Fuel Cycles, Nimes, France, June 21-24, 2004; 012- 01.

5. Weaver, B.; Kappelmann, F. A. Talspeak, A new method of separating americium and curium from the lanthanides by extraction from an aqueous solution of an aminopolyacetic acid complex with a monoacetic organophosphate or phosphonate. August 1964, ORNL-3559.

6. Nilsson, M.; Nash, K. L. Review Article: A Review of the Development and Operational Characteristics of the TALSPEAK Process. Solv. Extr. Ion Exch., 2007, 25, 665.

7. Kolarik, Z.; Koch, G.; Kuhn, W. Acidic Organophosphorus Extractants-XVIII The Rate of Lanthanide(III) Extraction by Di(2-ethylhexyl) Phosphoric Acid From Complexing Media. J. Inorg. Nucl. Chem. 1974, 36, 905.

8. Nilsson, M.; Nash, K. L. Manuscript in preparation.

9. Nash, K. L.; Brigham, D.; Shehee, T. C.; Martin, A. The kinetics of lanthanide complexation by EDTA and DTPA in lactate media. Dalton Trans. 2012, 41, 14547.

12

Chapter 2

Experimental

The rates of complex formation and dissociation of f-elements and polyaminopolycarboxylate (PAPC) ligands have been investigated previously primarily using either metal ion or ligand exchange reactions. A variety of techniques have been employed including spectophotometry,1-3 NMR spectroscopy,4-7 electrochemistry,8 and radiometric methods.9-12 The exchange reactions are mainly used for determining dissociation rates as the complex formation reactions for f-elements and PAPC ligands are generally very rapid. Direct measure of complex formation reactions have been reported (in dilute acetate buffer media) for only trivalent La2 and trivalent Am3 complexation by trans-1,2-diaminocyclohexane tetraacetic acid (DCTA). In these cases the reaction was slow enough to monitor directly due to the structural rigidity of the cyclohexyl backbone of DCTA, which hinders the conformational change of ligand that must take place in order to bind at all of its sites to the metal. It is interesting that the prearrangement of ligand donor atoms in what should be a favorable conformation (complex stability is increased about 100-fold for DCTA relative to EDTA) results in slower kinetics of complexation.

Spectrophotometric experiments for measuring the kinetics of rapid reactions often employ stopped-flow techniques. In a stopped-flow experiment, two syringes, each containing one component of the reaction, are set up parallel with a single driving piston on the plungers.

When signaled to fire, the driving piston acts on the syringes causing equal volumes of the reagent solutions to flow into a mixing chamber then onto an observation cell. The cessation of solution flow triggers the electronics to acquire an optical signal that is correlated with the

13 chemical reaction of interest. This method allows for the measurement of reactions with half- lives on the scale of milliseconds. In the series of experiments to follow the direct measurement of the complexation reaction between the lanthanides and PAPC ligands is achieved by employing stopped-flow spectrophotometry using a technique reported previously that makes these reactions kinetically accessible. As these reactions are otherwise too fast to be measured with such techniques, this result is primarily ascribed to the unusual aqueous media used in these experiments of high total concentrations of lactate, the effects of which will be discussed in detail in subsequent chapters.

The experimental method of equilibrium perturbation by ligand displacement that is applied in these series of experiments has been demonstrated in previous studies of examining different classes of actinide-complexing ligands.13-17 The lanthanides do not typically form strongly colored complexes or demonstrate significant changes in their absorption spectra from complexation with the PAPC ligands of interest in this study. Therefore, a strongly colored auxiliary reagent is required for spectrophotometric techniques such as stopped-flow to monitor the growth of the Ln-PAPC complexes. The colorimetric indicator dye 2,2’-(1,8-dihydroxy-3,6- disulfo-2,7-napthalene-bis(azo)dibenzenarsonic acid, known as Arsenazo III (hereafter AAIII)

(Figure 2.1) is used in these experiments for monitoring the rate of complexation of the lanthanides and PAPC ligands. It will be established that AAIII plays no significant role in the

Ln-PAPC complexation kinetics.

14

Figure 2.1: Structure of the Arsenazo III.

Free AAIII is magenta in color with a broad, intense absorption centering around 550 nm.

It has eight ionizable protons, for which acid dissociation constants have been reported (pK1 =

−2.5, pK2 = 0, pK3 = pK4 = 2.5, pK5 and pK6 = 5.3, pK7 = 7.5 and pK8 = 12.4); the most basic hydrogen ions are likely associated with the phenolic groups, the weakest with the sulfonate groups. AAIII complexes with the lanthanides are strongly colored blue with absorption bands at

610 and 658 nm. Their features have been investigated in some detail.18-22 The extinction

4 -1 -1 coefficients of the bands at the λmax= 658 nm are greater than 10 M cm , allowing the spectrophotometric monitoring of lanthanide concentration at 10-4 M or less, a critical design element for this investigation. It is found that AAIII forms 1:1 complexes with the lanthanides under most conditions and is believed to coordinate the lanthanide cation in a tetradentate manner with one oxygen atom from each arsenate and the phenol groups. AAIII is especially useful in analytical applications as it will complex with the lanthanides even under acidic conditions.

It has been demonstrated in previous studies that AAIII and Chlorophosphonazo III

(CLIII) interact with tetra,penta- and hexavalent actinide ions at a stopped-flow accessible

15 rate.23−27 However, studies looking at the interaction of AAIII with lanthanide ions found the complexation to mainly occur at or near the diffusion-controlled limit, too fast for monitoring with stopped-flow under normal conditions.28 This rapid interaction is advantageous in that the

Ln-AAIII complex rapidly exchanges between the complexed and free species, which allows for the PAPC ligand to complex with the free lanthanide. The subsequent decrease in the Ln-AAIII absorption spectrum can then be correlated to the growth of the Ln-PAPC complex, allowing for direct measurement of the forward complexation reactions between the lanthanides and PAPC ligands. Figure 2.2 shows the spectral change of the Eu-AAIII complex when mixed with DTPA.

0.6

0.4

Abs

0.2

0.0 560 580 600 620 640 660 680 700 Wavelength (nm)

Figure 2.2: Change in the visible spectrum upon mixing of EuAAIII with DTPA. 0.05 mM Eu 0.0125 mM AAIII and 8 mM DTPA in 1.0 M NaLac, pH 2.6, 2 M ionic strength 25 °C.

16

It should be noted that in the high lactate media employed in this series of experiments the Ln-

AAIII absorption band at 610 nm overlaps with the free AAIII spectra, however the 610 peak is found to be present after deconvolution of the spectra.

In the experiments that are to follow, an otherwise significantly complex system is simplified by operating under pseudo-first order conditions to monitor the approach to equilibrium between the lanthanides and the PAPC ligands. Under pseudo-first order conditions the concentration of the ligand is many times greater than that of the lanthanide ion, such that the concentration of the ligand is essentially constant over the course of the reaction. This allows for simple first-order analysis of the kinetic data of an otherwise second order system, yielding a first-order observed rate constant (kobs). The value of kobs will be dependent on the concentration of the ligand. This pseudo-first order approach to equilibrium method of kinetic analysis is a very powerful tool for investigating higher order reactions. It is instructive to first consider the fundamental mathematics of the kinetics of a first order approach to equilibrium reaction, as follows.

In a first order reversible reaction where reactant A goes to product P, the approach to equilibrium can be described in the rate of disappearance of the reactant. The system can be described as follows:

k 1 (1) k-1

(2)

(3)

(4)

17

Combining equations 2, 3, and 4 gives eq.5. Manipulation of the variables gives eq.6, and simple integration leads to eq.7.

(5)

(6)

(7)

It is seen here that a plot of ln([A]t - [A]e) with time will be linear, the negative of the slope of which is the observed rate constant of the reaction and is the sum of the forward and reverse rate constants.

This same treatment can be applied to a second order reversible approach to equilibrium reaction where reactants A and B go to product P, given the pseudo-first order condition of

[B]o >> [A]o. The system can be described as follows:

k 1 (8) k-1

(9)

(10)

(11)

(12)

Combination of equations 9, 10, and 11 and application of the pseudo-first order condition described in eq. 12 give eq. 13, manipulation of the variables and integration gives eq. 14.

18

(13)

(14)

Again it is seen here that a plot of ln([A]t - [A]e) with time will be linear, the negative slope of which is the observed rate constant of the pseudo-first order reaction:

(15)

This result shows that a plot of kobs against [B]o will give a straight line, the slope of which is the second-order rate constant of the forward reaction and the intercept is the first order rate constant of the reverse reaction. In the experiments that are to follow k1 will be the second-order rate

-1 -1 constant of complex formation kf (M s ) and k-1 will be the first-order rate constant of complex

-1 dissociation kd (s ). Equation 16 reflects the analysis of all the kinetic experiments to be presented.

(16)

Reagents

Lanthanide perchlorate stock solutions have been prepared, using 99.999% Ln2O3 from

Arris International Corp., through dissolution of the solid in HClO4. Stock solutions were standardized by ICP-OES analysis and sample acidity determined by potentiometric titration.

These lab stocks were used to prepare each lanthanide sample in all experiments. AAIII stock solutions were prepared by weight using material obtained from Alfa Aesar. Stock solutions of

EDTA were prepared by weight using 99.9% disodium EDTA salt from J.T. Baker. Stock solutions of DTPA and HEDTA were prepared by weight from the solid acid form and were

19 dissolved in two mole equivalents of sodium hydroxide to help with dissolution. The DTPA and

HEDTA solids were obtained from Alfa Aesar and Fluka Chemika respectively. Both were assayed at greater than 98%. Sodium lactate stock solutions were prepared by mass using 60% by mass sodium DL-lactate from Alfa Aesar. Solutions were acidified using stock solutions of perchloric acid prepared with 70% HClO4 from Fisher Scientific, and were standardized with

NaOH by indicator titration. NaOH solutions were prepared by mass using 50% NaOH from

Sigma Aldrich and standardized by KHP titration. The ionic strength of all solutions was adjusted to 2 M using a re-crystallized NaClO4 solution prepared in house.

Procedure

The kinetics of the reaction of the lanthanides with PAPC chelating agents in lactate media were investigated by stopped-flow spectrophotometry. Samples were thermally equilibrated for at least five minutes in the thermostated water bath of the stopped-flow unit prior to mixing. All kinetic experiments were performed using an OLIS U.S.A. stopped-flow system with a 20 mm optical cell paired with an OLIS RSM-1000 scanning spectrophotometer, except for experiments where the final DTPA concentration after mixing was greater than 6 mM. In this case, the stopped-flow system was paired with an OLIS Cary-14 spectrophotometer. The RSM-

1000 monitors a 220 nm wide window of wavelengths at speeds up to 1000 scans per second, and in all experiments this window was centered at 660 nm. The Cary-14 monitors a single wavelength and therefore can obtain a large number of data points for that wavelength in a short period of time. As such the Cary-14 was employed in experiments where the reaction reached completion in under 0.1 seconds. The Cary-14 was set to monitor at 660 nm in all experiments.

Kinetic data from the RSM-1000 was analyzed using the OLIS RSM Robust Global Fitting software package and observed rate constants were fit using first-order parameters. Kinetic data

20 from the Cary-14 was exported to Excel and observed rate constants were then fit in Origin 8 using a first-order decay model.

Trans-lanthanide Study

The complexation reaction of the ligands DTPA, EDTA, and HEDTA in high concentration lactate media was investigated with the lanthanides Pr3+, Nd3+, Sm3+-Lu3+ at

25.0 °C. In all cases the samples were made in solutions of 1.0 M NaLac, with the total ionic strength of the solution adjusted to 2.0 M with NaClO4. The lanthanide series was investigated under these conditions with each of the ligands at three different total added acid contents of 0.9,

0.85, and 0.8 M HClO4. At equilibrium these solutions have pH values of approximately 2.6, 2.8, and 3.0 respectively. The lanthanide solutions were made at 0.1 mM in all cases. The lanthanides were found to have a decreasing interaction with AAIII going across the series, thus it was necessary to increase the concentration of AAIII with heavier lanthanides. The AAIII concentration was 0.025 mM for Pr3+-Gd3+, 0.05 mM for Tb3+-Ho3+, 0.1 mM for Er3+-Yb3+ and

0.2 mM for Lu3+. The concentrations of DTPA and EDTA were made up to 4, 6, 8, 10 and 12 mM for reaction with the lanthanides. As the HEDTA reactions were found to be considerably slower, solutions were made at concentrations of 10, 20, 30, 40, and 50 mM in order to extrapolate a trend out of the change in the observed rate constants. For comparison, experiments with Eu3+ and higher concentrations of DTPA and EDTA were performed at 20, 30, 40, 50, and

60 mM at pH values of 2.6, 2.8, 3.0 with DTPA and pH 2.6 with EDTA. In all experiments equal volumes of the reacting solutions were mixed separately for pH measurement. The pH values were determined via pH probe calibrated with 0.01 and 0.001 M HClO4 standards made up in

2.0 M NaClO4.

21

Temperature Studies

In order to obtain activation parameters associated with the complex formation reaction, temperature studies were performed with the lanthanides Nd3+, Eu3+, Tb3+, Ho3+, Tm3+, and Lu3+ in 1.0 M NaLac, pH 2.6, and 2 M ionic strength with each of the ligands investigated at the same concentrations used in the trans-lanthanide study. As the rates were found to be significantly different between the different ligands and across the lanthanide series different temperature ranges were required for several of the different systems. In the DTPA temperature studies of

Nd3+ were run at 17.0, 20.0, 23.0 and 25.0 °C; Eu3+ at 10.0, 15.0, 20.0 and 25.0 °C; Tb3+ at 17.0,

20.0, 23.0, and 25.0 °C; Ho3+ at 20.0, 23.0, 25.0, 30.0, and 33.0 °C; Tm at 25.0, 27.0, 30.0, and

33.0 °C; Lu3+ at 25.0, 30.0, 35.0 and 40.0 °C. In the EDTA system, temperature studies of Nd3+ were run at 20.0, 25.0, 27.0, and 30.0 °C; Eu3+ and Tb3+ were run at 15.0, 20.0, 25.0, and

30.0 °C; Ho3+, Tm3+, and Lu3+ were run at 25.0, 27.0, 30.0, and 33.0 °C. In the HEDTA temperature studies the same range was used for all the lanthanides investigated at 25.0, 27.0,

30.0, and 33.0 °C. Reported temperatures have uncertainty of ±0.1 °C. In all cases the second- order rate constants were used in linear least squares fit Arrhenius plots and treated using

Activated Complex Theory (ACT) to obtain the activation parameters of the complex formation reaction.

Total Lactate Experiments

Often in kinetic experiments the role of one species in the reaction will be determined by maintaining constant concentrations of all the other reagents and varying the concentration of the species of interest. However in the high lactate system this is not possible as the concentration of lactic acid, lactate ion and hydrogen ion are all coupled. Since the concentration of one of these

22 species cannot be changed without changing the concentration of the others, a series of experiments were designed to determine the role of each of these species. Experiments examining the complexation kinetics between Eu3+ and DTPA were done under three different total lactate conditions of 1.0, 0.9, and 0.8 M. Each of these [Lac]tot systems were adjusted to have either the same lactate ion concentration or the same pH. This was done by calibrating a pH electrode under 2.0 M ionic strength conditions via a Gran titration. The results of the electrode calibration where then used to calculate the requisite mV reading for the desired pH in each system. Solutions were prepared and then adjusted with HClO4 or NaOH to within 1 mV of the calculated value. Solutions of 0.1 mM Eu3+ 0.025mM AAIII were mixed with DTPA concentrations of 4, 6, 8, 10, 12, 20, 30, 40, 50, and 60 mM under each [Lac]tot system adjusted to [Lac-] of 0.1, 0.2, and 0.3 M and the same pH of 2.72, 3.07, and 3.30.

AAIII Independence

In this series of experiments the kinetics of complexation between the lanthanides and

PAPC ligands is being measured by the perturbation of the equilibrium of the Ln-AAIII complex. The complexation reaction between hydrated lanthanides and AAIII has been reported as too fast to monitor by stopped-flow techniques at 25 °C.28 Therefore these experiments were operated with the assumption that the reaction mechanism was independent of the concentration of AAIII. However it is still prudent to test the system for a dependence on AAIII. Figure 2.3 shows the complexation of 0.05 mM Gd with EDTA in 1 M NaLac at 0.9 M HClO4 at two concentrations of AAIII.

It is seen in figure 2.3 that doubling the AAIII concentration has no effect on the kf value

-1 -1 (~2800 M s ), and the difference in the kd value can be attributed to the slight difference in the

23 pH as an acid dependant dissociation pathway for the lanthanides and PAPC ligands has been observed previously.29 These observations suggest that AAIII is not involved in the complexation reaction in these systems and therefore allows for the perturbation of the Ln-AAIII equilibrium as a method for monitoring the forward reaction of the lanthanides with the PAPC ligands.

22

20

18

16

14

) 12

-1

(s 10 obs k 8

6

4

2

0 0.000 0.001 0.002 0.003 0.004 0.005 0.006 EDTA (M)

Figure 2.3: Complexation of 0.05 mM Gd and EDTA 1.0 M NaLac 0.9M HClO4 2 M ionic strength 25 °C with [AAIII] (■) 0.0125 mM and (●) 0.025 mM at pH 2.7 and 2.8 respectively.

24

References

1. Nash, K. L.; Choppin, G. R., Sullivan, J. C. A Kinetic Study of Americium(III) and trans- 1,2 Diaminocyclohexane Tetraacetate. Inorg. Chem.1978, 17, 3374.

2. Nyssen, G. A.; Margerum, D. W. Multidentate Ligand Kinetics. XIV. Formation andDissociation Kinetics of Rare Earth-Cyclohexylenediaminetetraacetate Complexes. Inorg.Chem. 1970, 9, 1814.

3. Ryhl, T. Kinetics studies of lanthanoid carboxylate complexes. I. The dissociation rates of Ytterbium, Lanthanum and Copper EDTA complexes. Acta Chem. Scand. 1972, 26, 3955.

4. Siddall, T. H.; Stewart, W. E. Proton magnetic resonance studies of ethylenediaminetetraacetate complexes of lanthanides. Inorg. Nucl. Chem. Letters 1969, 5, 421.

5. Ryhl, T. Kinetics studies of lanthanoid carboxylate complexes. II. PMR investigation of the lanthanum and lutetium EDTA complexes. Acta Chem. Scand. 1972, 26, 4001.

6. Alsaadi, B. M.; Rossotti, F. J. C.; Williams, R. J. P. Hydration of complexon complexes of lanthanide cations. J. Chem Soc. Dalton Trans. 1980, 2151.

7. Gennaro, M. C.; Mirti, P., Casalino, C. NMR study of intramolecular processes in EDTA metalcomplexes. Polyhedron 1983, 2, 13.

8. Ryhl, T. Kinetics studies of lanthanoid carboxylate complexes. III. The dissociation rates of Praseodymium, Neodymium, Europium, and Erbium EDTA complexes. Acta Chem. Scand. 1973 27, 303.

9. Choppin, G. R.; Williams, K. R. Kinetics of Exchange Between Americium (III) and Europium Ethylenediaminetetraacetate. J. Inorg. Nucl. Chem. 1973, 35(12), 4255.

10. Williams K. R.; Choppin, G. R. J. Kinetics of exchange of trivalent actinide ions with europium ethylenediaminetetraacetate. Inorg. Nucl. Chem. 1974, 36, 1849.

11. Muscatello, A. C.; Choppin, G. R.; D'Olieslager, W. Kinetics of dissociation of trivalent actinide chelates of trimethylenediamine-N,N,N',N'-tetraacetic acid [TMDTA]. Inorg. Chem. 1989, 28(6), 993.

12. Glentworth, P.; Wiseall, B.; Wright, C. L.; Mahmood, A. J. A Kinetic Study of Isotopic Exchange Reactions Between Lanthanide Ions and Lanthanide Polyaminopolycarboxylic Acid Complex Ions. I. Isotopic Exchange Reactions of Ce(III) with Ce(HEDTA), Ce(EDTA)-, Ce(DCTA)-, and Ce(DTPA)2-. J. Inorg. Nucl. Chem. 1968, 30(4), 967.

25

13. Hines, M. A.; Sullivan, J. C.; Nash, K. L. Study of the Complexation Kinetics of Dioxouranium(VI) with Selected Diphosphonic Acids in Acidic Solutions. Inorg. Chem. 1993, 32, 1820.

14. Fugate, G. A.; Nash, K. L.; Sullivan, J. C. Kinetic Study of the Reactions of Np(V) and U(VI) with Diphosphonic Acids in Acetate Buffer Solutions. Radiochim. Acta 1997 79, 161.

15. Friese, J. I.; Nash, K. L.; Jensen, M. P.; Sullivan, J. C. Kinetic Study of the Reactions of Np(V) and U(VI) with Oxydiacetic Acid. Radiochim. Acta 1998, 83, 175.

16. Friese, J. I.; Nash, K. L.; Jensen, M. P.; Sullivan, J. C. Interaction of Np(V) and U(VI) with Dipicolinic Acid. Radiochim. Acta 2001, 89, 35

17. Hall, H.; Sullivan, J.C.; Rickert, P.G.; Nash, K.L. Kinetics of the Reaction of U(VI ) With Benzene-1,2-Diphosphonic Acid. Dalton Trans. 2005, 2011.

18. Rowatt, E.; Williams, R. The Interaction of Cations with the Dye Arsenazo III. Biochem.J. 1989, 259, 295.

19. Rohwer, H.; Collier, N; Hosten, E. Spectrophotometric study of arsenazo III and its interactions with lanthanides. Anal. Chim. Acta 1995, 314, 219.

20. Rohwer, H.; Hosten, E. pH dependence of the reactions of arsenazo III with the lanthanides. Anal. Chim. Acta 1997, 339, 271.

21. Hosten, E.; Rohwer, H. Interaction of anions with arsenazo III-lanthanide (III) complexes. 1997, 345, 227.

22. Lu, Y. W.; Laurent, G.; Pereira, H. A novel methodology for evaluation of formation constants of complexes: example of lanthanide-Arsenazo III complexes. Talanta 2004, 62, 959.

23. Pippin, C. G.; Sullivan, J. C.; Wester, D. W. A kinetic study of the reactions between tetravalent actinide ions and Arsenazo(III). Radiochim. Acta 1984, 37, 99.

24. Pippin, C. G.; Sullivan, J. C. A kinetic study of the complexation reaction between the dioxouranium(VI) ion and Arsenazo III. Radiochim. Acta 1989, 48(1-2), 37.

25. Feil Jenkins, J. F.; Sullivan, J. C.; Nash, K. L. Kinetics of the Complexation of Dioxouranium(VI) with Chlorophosphonazo III. Radiochim. Acta 1995, 68, 209.

26

26. Feil Jenkins, J.; Nash, K. L.; Sullivan, J. C. Kinetics of the Complexation Reactions Between Th(IV), Zr(IV) and Chlorophosphonazo III. Radiochim. Acta 1995, 68, 215.

27. Fugate, G.; Feil-Jenkins, J. F.; Sullivan, J. C.; Nash, K. L. Actinide Complexation Kinetics: Rate and Mechanism of Dioxoneptunium(V) Reaction with Chlorophosphonazo III. Radiochim. Acta 1996, 73, 67.

28. Shi, Y; Eyring, E; van Eldik, R. Kinetic analysis of the complexation of aqueous lanthanides(III) ions by arsenazo III. Dalton Trans. 1998, 6, 967.

29. Choppin, G. Complexation kinetics of f-element and polydentate ligands. Journal of Alloys and Compounds 1995, 225, 242.

27

Chapter 3

Experimental Results

As mentioned in the previous chapter, the rates of complex formation and dissociation of f-elements and PAPC ligands have been examined using a variety of techniques.1-12 The studies investigating Ln-PAPC reactions looked at complex dissociation rates as the formation reactions were too fast to monitor. Efforts to calculate formation rate constants between lanthanides and

EDTA from dissociation rate data by Laurenczy and Brücher13 found rate constants on the order of 106 M-1s-1 and 108 M-1s-1 for diprotonated and monoprotonated EDTA respectively. They determined a formation rate constant for monoprotaonated EDTA with Gd3+ of 3.0 x 108 M-1s-1 which is comparable to, but slightly slower than, the reported value for Gd3+ water exchange at

10.6 x 108 s-1.14 Additionally, values for the rate constants of complexation between the lanthanides and DCTA, calculated by Nyssen and Margerum, are found to be rapid despite the conformational hindrance of the cyclohexyl-backbone, with values on the order of 107 M-1s-1.2

These consistently rapid rates of reaction suggest that the complexaiton of lanthanides with

PAPC ligands takes place at or near the diffusion limit under normal aqueous conditions. It is notable then that in the experimental results to follow, the complexation of the lanthanides with the ligands DTPA, EDTA and HEDTA are directly measured. Being able to directly monitor these reactions speaks to the considerable effect that the high concentration of lactate has on these systems. Reactions that are normally on a time scale accessible only by NMR are slowed to the point of being able to be monitored by stopped-flow spectrophotometry. The experiments presented here aim to elucidate the significant effect the lactate has on these systems. Errors associated with all reported values correspond to a 95% confidence interval.

28

DTPA

Lanthanide Series Study

The Eu-DTPA complexation results shown in figure 3.1 are representative for the

lanthanide series. Table 3.1 shows the second order rate constant of complex formation and table

3.2 shows the first order rate constant of complex dissociation between the lanthanides and

DTPA in 1.0 M NaLac for the three acidities studied at 2.0 M ionic strength and 25.0 °C. It is

seen that across the lanthanides studied there is a trend of decreasing rate constants for both

formation and dissociation with decreasing acidity. This pattern is strictly adhered to for the kd

values, while the kf values mainly follow the trend. For lanthanides where the trend in kf values is

80

70

60

50 )

-1 40

(s obs

k 30

20

10

0 0.000 0.001 0.002 0.003 0.004 0.005 0.006 DTPA (M)

Figure 3.1: Observed rate constants of the reaction between Eu3+ and DTPA in 1.0 M NaLac 0.9 (■), 0.85 (●), and 0.8 () M HClO4 2 M ionic strength at 25 °C.

29

-1 -1 kf (M s )

Ln 0.9 M HClO4 0.85 M HClO4 0.8 M HClO4 Pr 6900 (114) 7170 (370) 6880 (150) Nd 7020 (145) 7200 (10) 6730 (200) Sm 10410 (190) 10070 (270) 9080 (110) Eu 11310 (90) 10430 (80) 9790 (90) Gd 9620 (20) 9540 (110) 8520 (220) Tb 7370 (50) 6670 (190) 5810 (60) Dy 5300 (80) 5410 (40) 3910 (50) Ho 2670 (20) 2780 (20) 2590 (20) Er 1820 (60) 1530 (20) 1080 (20) Tm 1420 (40) 980 (30) 670 (100) Yb 950 (30) 670 (30) 530 (30) Lu 470 (10) 500 (10) 560 (20) Table 3.1: Lanthanide-DTPA second order rate constant of complex

formation. 1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C.

-1 kd (s )

Ln 0.9 M HClO4 0.85 M HClO4 0.8 M HClO4 Pr 21.0 (0.6) 16.7 (1.1) 15.1 (0.7) Nd 11.0 (0.3) 9.4 (0.1) 8.6 (0.6) Sm 7.4 (0.8) 5.3 (0.8) 4.5 (0.2) Eu 5.4 (0.2) 3.4 (0.4) 2.1 (0.4) Gd 5.0 (0.1) 1.4 (0.2) 0.8 (0.5) Tb 4.6 (0.1) 2.2 (0.5) 1.3 (0.2) Dy 3.1 (0.3) 1.6 (0.1) 0.6 (0.3) Ho 2.7 (0.1) 1.2 (0.1) -0.1 (0.1) Er 2.8 (0.1) 1.4 (0.1) 1.2 (0.1) Tm 2.7 (0.2) 1.4 (0.1) 1.0 (0.4) Yb 2.8 (0.1) 1.6 (0.1) 0.8 (0.1) Lu 3.3 (0.1) 1.5 (0.1) 0.1 (0.1) Table 3.2: Lanthanide-DTPA first order rate constant of complex dissociation.

1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C. not followed, the kf values at the different acidities are close to one another i.e. Dy at 0.9 M

-1 -1 -1 -1 HClO4 and 0.85 M HClO4 with kf values of 5300 ± 80 M s and 5410 ± 80 M s respectively.

30

Ln-DTPA complex formation and dissociation rate constants are plotted against the inverse ionic radii of the lanthanides in figures 3.2 and 3.3, respectively. It is seen in figure 3.2 that the second order rate constant of complex formation is essentially constant for Pr3+ and Nd3+

3+ 3+ at all acidities, followed by a large increase in kf at Sm which peaks at Eu . This sudden increase in the rate of complex formation may be attributed at least in part to the change in the

12000

10000

8000

)

-1

s -1

6000

(M

f k 4000

2000

0 0.84 0.86 0.88 0.90 0.92 0.94 0.96 1/r (CN=9, Angstoms-1)

Figure 3.2: Ln-DTPA kf values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M

HClO4, 2 M ionic strength at 25 °C. coordination environment of the lanthanides that occurs midway through the series due to the decreasing ionic radii.15 Sm3+ and Eu3+ are in equilibrium between species of coordination number (CN) of 9 and 8, and those ions with a CN of 9 will more easily lose a coordinated water allowing DTPA to more readily bind with the metal. The steady and significant decrease in kf that occurs after Eu3+ would not be expected considering the decrease in the ionic radii that

31

25

20

15

)

-1

(s d

k 10

5

0 0.84 0.86 0.88 0.90 0.92 0.94 0.96 1/r (CN=9, Angstroms-1)

Figure 3.3: Ln-DTPA kd values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M

HClO4, 2 M ionic strength at 25 °C. occurs across the lanthanide series, which leads to an increase in charge density. With an increase in charge density it is reasonable to expect an increase in the interaction between the trivalent lanthanide ion and the negatively charged carboxylate groups on the DTPA leading to faster coordination. This behavior could be expected from the interaction between DTPA and hydrated lanthanide ions, however given the high concentration of lactate in the system, starting from the hydrated lanthanide ions may not be appropriate. Taking the lanthanide-lactate species into consideration for their role in the kinetics will be shown to have a significant effect. This argument will be developed in great detail in the discussion chapter.

32

In contrast to the behavior of the kf values, the trend of decreasing kd values that is observed going across the lanthanide series is behavior that would be expected with the decrease in ionic radii. The resulting increase in charge density leads to stronger interaction with the binding groups of DTPA thus the lanthanide is less likely to dissociate. Additionally, the observed trend of decreasing kd with decreasing acidity would be expected. An acid catalyzed pathway has been observed regularly in the dissociation of the lanthanides and many PAPC ligands, thus in systems with lower hydrogen ion concentrations slower rates of dissociation are should be expected.16

Temperature Study

To calculate activation parameters for the complexation reaction between DTPA and the lanthanides Nd3+, Eu3+, Tb3+, Ho3+, Tm3+, and Lu3+ temperature studies were completed. As the rates of complexation for these lanthanides cover a wide margin, different temperature ranges were required as described in the experimental chapter. Figure 3.4 shows the reaction of Eu3+ with DTPA at the temperatures of 25.0, 20.0, 15.0, and 10.0 °C and is representative for the lanthanides studied. Table 3.3 lists the activation parameters for the Ln-DTPA complex formation reaction in 1.0 M NaLac and 0.9 M HClO4 at 2 M ionic strength. It is seen for all the lanthanides studied that the activation entropy (ΔS‡) is highly positive, which is indicative of a dissociative reaction mechanism. Surprisingly, the activation enthalpy (ΔH‡) for these reactions are all very similar and do not follow the expected trend of lower activation energies for faster rates of reaction. This surprising result is well illustrated by comparing the activation enthalpies and the rate constants of complex formation at 25 °C of Eu3+ and Lu3+ which are ΔH‡ = 96 ± 3

-1 -1 ‡ -1 -1 kJ/mol, kf = 11300 ± 90 M s and ΔH = 107 ± 4 kJ/mol, kf = 470 ± 10 M s respectively. The rate of complexation for Lu3+ is 24 times slower than that of Eu3+ however the activation

33 parameters are comparatively close to one another, as is the case for all the lanthanides studied.

The similar activation parameters across the lanthanide series may be suggestive of a very similar mechanism for complex formation between the different metals, despite the significant difference in the rates of reaction across the series.

80

60 )

-1 40

(s

obs k

20

0 0.000 0.001 0.002 0.003 0.004 0.005 0.006 DTPA (M)

3+ Figure 3.4: Temperature dependence of the reaction between Eu and DTPA 1 M NaLac 0.9 M HClO4 2 M ionic strength at 25 °C (■), 20 °C (●), 15 °C () and 10 °C ().

34

E ΔG‡ ΔH‡ ΔS‡ k a ln(A) f Ln kJ/mol kJ/mol kJ/mol J/mol*K 25 °C M-1s-1 Nd 97 (4) 48 (2) 52 (5) 95 (4) 146 (13) 7020 (145)

Eu 99 (3) 49 (1) 51 (4) 96 (3) 156 (10) 11300 (90)

Tb 100 (2) 49 (1) 51 (3) 97 (2) 155 (8) 7370 (50) Ho 90 (2) 44 (1) 53 (7) 88 (2) 117 (5) 2870 (20) Tm 83 (6) 42 (2) 55 (9) 80 (6) 85 (21) 1420 (40) Lu 109 (4) 50 (2) 56 (6) 107 (4) 164 (15) 470 (10) Table 3.3: Activation parameters for the Ln-DTPA complexation reaction

in 1.0 M NaLac 0.9 M HClO4 and 2 M ionic strength. Additionally the second order rate constant of complex formation at 25 °C is listed.

EDTA

Lanthanide Series Study

The Eu-EDTA complexation results shown in figure 3.5 are representative for the lanthanide series. Table 3.4 shows the second order rate constant of complex formation and table

3.5 shows the first order rate constant of complex dissociation between the lanthanides and

EDTA in 1.0 M NaLac for the three acidities studied at 2.0 M ionic strength and 25.0 °C. As with DTPA, it is seen that across the lanthanides studied there is a trend of decreasing rate constants for both formation and dissociation with decreasing acidity. This pattern is strictly

3+ adhered to for the kf values, except for at Ho where at 0.85 and 0.8 M HClO4 the kf values are within error of each other. The trend with acidity is followed in the kd values for the early and mid lanthanides, which indicates an acid catalyzed dissociation pathway as has been observed

35

22

20

18

16

14

) 12

-1

(s 10 obs k 8

6

4

2

0 0.000 0.001 0.002 0.003 0.004 0.005 0.006 EDTA (M)

Figure 3.5: Observed rate constants of the reaction between Eu3+ and EDTA in 1.0 M NaLac and 0.9 (■), 0.85 (●), and 0.8 () M HClO4 2 M ionic strength at 25 °C.

previously in PAPC ligands.16 At the heavy lanthanides the rate of dissociation becomes very

slow and the values at all acidities are within error of each other. While this may indicate the

acid dependency falls off at the end of the series care must be taken in interpreting these results

as some of the values are not statistically different from zero. The dissociation rates with the

heaviest lanthanides may be too slow to reliably measure with this extrapolative process.

Ln-EDTA complex formation and dissociation rate constants are plotted against the

inverse ionic radii of the lanthanides in figures 3.6 and 3.7, respectively. EDTA and DTPA

demonstrate similar behavior as it is seen in figure 3.6 that the second-order rate constant of

36

kf

Ln 0.9 M HClO4 0.85 M HClO4 0.8 M HClO4 Pr 1350 (20) 1250 (20) 1080 (20) Nd 1310 (30) 1180 (50) 1040 (10) Sm 2340 (50) 1910 (20) 1720 (20) Eu 2650 (10) 2310 (30) 2080 (30) Gd 2760 (50) 2410 (120) 2060 (40) Tb 2830 (50) 2220 (70) 2090 (30) Dy 2390 (60) 2060 (30) 1620 (50) Ho 1810 (20) 1010 (10) 1040 (20) Er 1160 (20) 880 (20) 670 (10) Tm 840 (50) 610 (10) 410 (10) Yb 500 (10) 320 (10) 290 (10) Lu 280 (10) 200 (10) 160 (10) Table 3.4: Lanthanide-EDTA second order rate constant of complex

formation. 1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C.

kd

Ln 0.9 M HClO4 0.85 M HClO4 0.8 M HClO4 Pr 7.7 (0.1) 5.6 (0.1) 4.3 (0.1) Nd 5.6 (0.2) 4.3 (0.1) 3.4 (0.1) Sm 4.2 (0.1) 3.4 (0.1) 2.9 (0.1) Eu 4.6 (0.1) 3.3 (0.1) 2.8 (0.1) Gd 4.3 (0.3) 3.1 (0.4) 2.3 (0.1) Tb 3.3 (0.1) 3.2 (0.2) 1.9 (0.1) Dy 2.3 (0.2) 1.4 (0.1) 1.6 (0.2) Ho 1.0 (0.1) 0.8 (0.1) 1.0 (0.1) Er 0.5 (0.1) 0.5 (0.1) 0.4 (0.1) Tm 0.4 (0.2) 0.2 (0.1) 0.3 (0.1) Yb 0.2 (0.1) 0.3 (0.1) 0.1 (0.1) Lu 0.2 (0.1) 0.2 (0.1) 0.1 (0.1) Table 3.5: Lanthanide-EDTA first order rate constant of complex dissociation.

1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C.

3+ 3+ complex formation is constant for Pr and Nd between the different acidities. An increase in kf is observed at all acidities beginning at Sm3+ which levels out at Eu3+ and the values do not change significantly until Dy3+ followed by a steady decrease for the rest of the series. The

37

3000

2500

2000

)

-1

s -1

1500

(M

f k 1000

500

0 0.84 0.86 0.88 0.90 0.92 0.94 0.96 1/r (CN=9, Angstroms-1)

Figure 3.6: Ln-EDTA kf values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M

HClO4, 2 M ionic strength at 25 °C.

3+ 3+ increase in kf that occurs between Nd and Eu can partially be attributed to the change in the coordination environment that occurs as the lanthanide ionic radius decreases across the series as

15 was seen for DTPA. Also like DTPA, the decrease in kf that occurs for the heavy lanthanides would not be expected considering the increase in charge density that occurs across the lanthanide series, which should increase the strength of the interaction between the metal ion and the carboxylate groups of EDTA. This is again indicative of more complex behavior in the lactate system than would be expected from the interaction between the ligand and hydrated lanthanide ions.

38

8

7

6

5 )

-1 4

(s d k 3

2

1

0

0.84 0.86 0.88 0.90 0.92 0.94 0.96 1/r (CN=9, Angstroms-1)

Figure 3.7: Ln-EDTA kd values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M

HClO4, 2 M ionic strength at 25 °C.

Temperature Study

Temperature studies were completed on the reaction between EDTA and the lanthanides

Nd3+, Eu3+, Tb3+, Ho3+, Tm3+, and Lu3+. Figure 3.7 shows the reaction of Eu3+ with EDTA at the temperatures of 30.0, 25.0, 20.0, and 15.0 °C and is representative for the lanthanides studied.

Table 3.6 lists the activation parameters for the Ln-EDTA complex formation reaction in 1.0 M

NaLac and 0.9 M HClO4 at 2.0 M ionic strength. As with DTPA, a dissociative reaction mechanism is indicated in all the lanthanides studied based on the highly positive ΔS‡ values.

39

Additionally both the ΔS‡ and ΔH‡ values for the different lanthanides are identical within the limits of uncertainty. As the kf values across the lanthanide series change considerably the similarity of the resulting activation parameters does not readily provide an explanation for the disparate values of the complex formation rate constants. There is an order of magnitude difference between the complex formation second order rate constant of Tb3+ and Lu3+, however the activation enthalpy values are 95 ± 3 kJ/mol and 96 ± 4 kJ/mol and the activation entropy values are 138 ± 10 J/(mol*K) and 134 ± 12 J/(mol*K) respectively. While these results do not explain the change in the kf values across the series, the similar activation parameters may suggest that the lanthanides complex to EDTA via a similar mechanism.

45

40

35

30

25

)

-1

(s 20

obs k 15

10

5

0 0.000 0.001 0.002 0.003 0.004 0.005 0.006 EDTA (M)

3+ Figure 3.8: Temperature dependence of the reaction between Eu and EDTA 1 M NaLac 0.9 M HClO4 2 M ionic strength at 30 °C (■), 25 °C (●), 20 °C () and 15 °C ().

40

E ΔG‡ ΔH‡ ΔS‡ k a ln(A) f Ln kJ/mol kJ/mol kJ/mol J/mol*K 25 °C M-1s-1 Nd 102 (11) 48 (4) 55 (15) 99 (11) 148 (35) 1310 (33) Eu 94 (2) 46 (1) 54 (3) 92 (2) 128 (7) 2650 (10) Tb 97 (3) 47 (1) 54 (4) 95 (3) 138 (10) 2830 (50) Ho 99 (2) 48 (1) 54 (3) 97 (2) 142 (7) 1810 (20) Tm 99 (4) 47 (1) 56 (5) 96 (4) 134 (12) 840 (50) Lu 98 (4) 46 (2) 56 (5) 96 (4) 126 (12) 280 (10) Table 3.6: Activation parameters for the Ln-EDTA complexation reaction in

1.0 M NaLac 0.9 M HClO4 and 2 M ionic strength. Additionally the second order rate constant of complex formation at 25 °C is listed.

HEDTA

Lanthanide Series Study

The Eu-HEDTA complexation results shown in figure 3.9 are representative for the lanthanide series. It is notable that the HEDTA concentrations used in these studies are considerably higher than those for the DTPA and EDTA systems. Under the experimental conditions the lanthanide complexation reaction with HEDTA is considerably slower and a larger range in the ligand concentration was required in order to see a substantial enough change in kobs to obtain adequate precision in the values of kf and kd.

41

5.0

4.5

4.0

3.5

3.0

) -1

2.5

(s obs

k 2.0

1.5

1.0

0.5

0.0 0.000 0.005 0.010 0.015 0.020 0.025 HEDTA (M)

Figure 3.9: Observed rate constants of the reaction between Eu3+ and HEDTA in 1.0 M NaLac and 0.9 (■), 0.85 (●), and 0.8 () M HClO4 2 M ionic strength at 25 °C.

Table 3.7 shows the second order rate constant for complex formation and table 3.8

shows the first order rate constant of complex dissociation between the lanthanides and HEDTA

in 1.0 M NaLac for the three acidities studied at 2.0 M ionic strength and 25.0 °C. In contrast to

the other two ligands studied, the kf values do not exhibit a trend with acidity. While the errors

on these values are fairly small in some cases, the scattering of similar values between the

different acidities for all the lanthanides suggests the complex formation reaction is independent

of acidity in the range investigated. The kd values display similar behavior except at the

3+ 3+ lanthanides Eu through Ho where the kd value for 0.9 M HClO4 are higher than for the two

lower acidities, the values of which fall within error of each other. This suggests that an acid

catalyzed dissociation pathway becomes important for the mid lanthanides at higher acidities (pH

42

< 2.6), that becomes less significant at lower acidity and at the heaviest of the lanthanides. The

kf

Ln 0.9 M HClO4 0.85 M HClO4 0.8 M HClO4 Pr 130 (8) 140 (6) 140 (8) Nd 120 (5) 110 (10) 120 (10) Sm 140 (3) 135 (1) 140 (1) Eu 130 (1) 135 (2) 140 (1) Gd 110 (2) 115 (1) 120 (2) Tb 110 (1) 105 (3) 105 (1) Dy 95 (1) 90 (1) 85 (2) Ho 70 (1) 80 (4) 65 (2) Er 65 (8) 60 (7) 65 (3) Tm 55 (11) 50 (3) 60 (3) Yb 55 (1) 45 (5) 40 (1) Lu 35 (7) 45 (5) 45 (8) Table 3.7: Lanthanide-HEDTA second order rate constant of complex

formation. 1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C.

kd

Ln 0.9 M HClO4 0.85 M HClO4 0.8 M HClO4 Pr 5.4 (0.2) 5.3 (0.1) 5.0 (0.1) Nd 3.4 (0.1) 3.5 (0.2) 3.4 (0.2) Sm 1.7 (0.1) 1.7 (0.1) 1.7 (0.1) Eu 1.8 (0.1) 1.4 (0.1) 1.4 (0.1) Gd 1.8 (0.1) 1.4 (0.1) 1.3 (0.1) Tb 2.1 (0.1) 1.7 (0.1) 1.6 (0.1) Dy 2.4 (0.1) 2.0 (0.1) 1.8 (0.1) Ho 3.0 (0.1) 2.1 (0.1) 2.2 (0.1) Er 2.5 (0.1) 2.4 (0.1) 2.5 (0.1) Tm 2.5 (0.1) 2.5 (0.1) 2.4 (0.1) Yb 1.4 (0.1) 1.5 (0.1) 1.5 (0.1) Lu 1.1 (0.1) 0.9 (0.1) 0.9 (0.2) Table 3.8: Lanthanide-HEDTA first order rate constant of complex

dissociation. 1.0 M NaLac variable HClO4 2.0 M NaClO4 ionic strength 25 °C.

43

lack of an apparent acid catalyzed dissociation pathway in this system is of interest, as it runs

counter to the observations made in the DTPA and EDTA systems, as well as what has been

reported in the literature.16

Ln-HEDTA complex formation and dissociation rate constants are plotted against the

inverse ionic radii of the lanthanides in figures 3.10 and 3.11, respectively. It is seen in figure

3.10 that the second order rate constant of complex formation stays essentially constant from

Pr3+ through Eu3+ and then steadily decreases through the rest of the series. The changing

coordination environment at the mid lanthanides does not appear to effect kf in the manner that

was seen for DTPA and EDTA. However the decrease in the rate of complex formation occurs at

the mid to late lanthanides is consistent with what was seen with the other ligands.

160

140

120

100

)

-1

s -1

80

(M

f k 60

40

20

0 0.84 0.86 0.88 0.90 0.92 0.94 0.96 1/r (CN=9, Angstroms-1)

Figure 3.10: Ln-HEDTA kf values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M HClO4, 2 M ionic strength at 25 °C.

44

6

5

4 )

-1 3

(s

d k 2

1

0 0.84 0.86 0.88 0.90 0.92 0.94 0.96 1/r (CN=9,Angstroms-1)

Figure 3.11: Ln-HEDTA kd values across the Ln series (Pr-Lu) in 1 M NaLac, 0.9 (■), 0.85 (●), and 0.8 () M HClO4, 2 M ionic strength at 25 °C.

Temperature Study

Temperature studies were completed to calculate activation parameters for the

complexation reaction between HEDTA and the lanthanides Nd3+, Eu3+, Tb3+, Ho3+, Tm3+, and

Lu3+. Figure 3.12 shows the reaction of Eu3+ with HEDTA at the temperatures of 33.0, 30.0,

27.0, and 25.0 °C and is representative for the lanthanides studied. Table 3.9 lists the activation

parameters for the Ln-HEDTA complex formation reaction in 1.0 M NaLac and 0.9 M HClO4 at

2.0 M ionic strength. As seen with the other ligands studied, the highly positive ΔS‡ values are

indicative of a dissociative reaction mechanism. Although the rates of complex formation

between the lanthanides and HEDTA are considerably slower in the lactate system than DTPA or

45

12

10

8

)

-1

(s 6

obs k

4

2

0 0.000 0.005 0.010 0.015 0.020 0.025 HEDTA (M)

3+ Figure 3.12: Temperature dependence of the reaction between Eu and HEDTA 1 M NaLac 0.9 M HClO4 2 M ionic strength at 33 °C (■), 33 °C (●), 27 °C () and 25 °C (). EDTA the same condition of similar activation parameters for different rates is present. Of the lanthanides investigated in the temperature study Eu3+ has the fastest rate constant of

-1 -1 3+ -1 -1 complexation at 130 ± 1 M s and Lu has the slowest kf value at 35 ± 7 M s and yet their

ΔH‡ and ΔS‡ are identical within the limits of uncertainty at 99 ± 2 kJ/mol and 126 ± 5 J/(mol*K) for Eu3+ and 105 ± 4 kJ/mol and 135 ± 14 J/(mol*K) for Lu3+. The similar activation parameters again suggest a similar mechanism for complex formation between the lanthanides and HEDTA in the lactate system. However as with the other ligands studied, the similar activation parameters do not provide an explanation for the decreasing kf values going across the lanthanide series. As this condition is seen in all the ligands studied, possible explanations for this behavior will be developed further in the discussion chapter.

46

‡ ‡ ‡ k Ea ΔG ΔH ΔS f ln(A) -1 -1 Ln kJ/mol kJ/mol kJ/mol J/mol*K 25 °C M s Nd 89 (3) 41 (1) 61 (5) 87 (3) 86 (10) 120 (5) Eu 101 (2) 46 (1) 61 (2) 99 (2) 126 (5) 130 (1) Tb 95 (3) 43 (1) 61 (4) 93 (3) 105 (9) 110 (1) Ho 93 (9) 42 (4) 53 (7) 91 (9) 96 (30) 70 (1) Tm 90 (5) 41 (2) 62 (7) 88 (5) 85 (17) 55 (11) Lu 107 (4) 47 (2) 64 (6) 105 (4) 135 (14) 35 (7) Table 3.9: Activation parameters for the Ln-HEDTA complexation reaction

in 1.0 M NaLac 0.9 M HClO4 and 2 M ionic strength. Additionally the second order rate constant of complex formation at 25 °C is listed.

High Ligand Concentration Study

It was seen in the HEDTA system that high concentrations of ligand were required to obtain values for kf and kd. A large increase in the concentration of any component of a reaction can potentially lead to a significant change in the reaction mechanism. To compare the HEDTA results with those of DTPA and EDTA, it is prudent to examine the DTPA and EDTA systems under the condition of high concentrations of the PAPC ligand. Figure 3.13 compares the results

3+ of the Eu complexation with DTPA, EDTA, and HEDTA in 1.0 M NaLac 0.9 M HClO4 and 2

M ionic strength at 25 °C.

47

In figure 3.13 it is seen that EDTA behaves similarly at high and low concentrations, however the behavior of DTPA changes going from low to high concentration realms. The increase in kobs with increasing [DTPA] begins to plateau at [DTPA]t > 15 mM which suggests a saturation effect at high concentrations of DTPA, in effect the plateau indicates that the reaction is becoming independent of [DTPA]t in the high concentration limit.

160

140

120

100 )

-1 80

(s obs k 60

40

20

0 0.000 0.005 0.010 0.015 0.020 0.025 0.030 [PAPC]

Figure 3.13: Complexation of Eu3+ with (■) DTPA, (●) EDTA, and () HEDTA in 1.0 M NaLac, 0.9 M HClO4, and 2 M ionic strength at 25 °C. Lines added as a visual guide.

48

Figure 3.14 shows the complexation of Eu3+ with the full range of DTPA concentrations at all acidities studied. It can be seen in figure 3.14 that kobs levels off at high [DTPA] at all the acidites studied, beginning at 15 mM DTPA. It is also observed that decreasing the acidity of the system decreases the value at which kobs plateaus. In a simple saturation system, it could be expected to see kobs level off at the same value for all acidities if the effect was dependant on the concentration of DTPA alone, however the behavior observed here suggests a more complex system.

160

140

120

100 )

-1 80

(s obs k 60

40

20

0 0.000 0.005 0.010 0.015 0.020 0.025 0.030 DTPA (M)

Figure 3.14: Observed rate constants of the reaction between Eu3+ and DTPA in 1.0 M NaLac 0.9

(■), 0.85 (●), and 0.8 () M HClO4 2 M ionic strength at 25 °C. Lines added as a visual guide.

49

It is worth reiterating that the preceding results probed reactions that have previously been characterized as too fast to monitor with stopped-flow spectrophotometry. The addition of lactate to the system allows for the complexation reactions to be directly monitored, however it also complicates the system considerably. This higher degree of complexity has yielded unexpected results in the trans-lanthanide studies as well as in the temperature studies. In the following chapter these results will be analyzed in greater detail to in an attempt to better understand the roles each of the different species has in the complexation kinetics of these complex systems.

50

References

1. Nash, K. L.; Choppin, G. R., Sullivan, J. C. A Kinetic Study of Americium(III) and trans- 1,2 Diaminocyclohexane Tetraacetate. Inorg. Chem.1978, 17, 3374.

2. Nyssen, G. A.; Margerum, D. W. Multidentate Ligand Kinetics. XIV. Formation andDissociation Kinetics of Rare Earth-Cyclohexylenediaminetetraacetate Complexes. Inorg.Chem. 1970, 9, 1814.

3. Ryhl, T. Kinetics studies of lanthanoid carboxylate complexes. I. The dissociation rates of Ytterbium, Lanthanum and Copper EDTA complexes. Acta Chem. Scand. 1972, 26, 3955.

4. Siddall, T. H.; Stewart, W. E. Proton magnetic resonance studies of ethylenediaminetetraacetate complexes of lanthanides. Inorg. Nucl. Chem. Letters 1969, 5, 421.

5. Ryhl, T. Kinetics studies of lanthanoid carboxylate complexes. II. PMR investigation of the lanthanum and lutetium EDTA complexes. Acta Chem. Scand. 1972, 26, 4001.

6. Alsaadi, B. M.; Rossotti, F. J. C.; Williams, R. J. P. Hydration of complexon complexes of lanthanide cations. J. Chem Soc. Dalton Trans. 1980, 2151.

7. Gennaro, M. C.; Mirti, P., Casalino, C. NMR study of intramolecular processes in EDTA metalcomplexes. Polyhedron 1983, 2, 13.

8. Ryhl, T. Kinetics studies of lanthanoid carboxylate complexes. III. The dissociation rates of Praseodymium, Neodymium, Europium, and Erbium EDTA complexes. Acta Chem. Scand. 1973 27, 303.

9. Choppin, G. R.; Williams, K. R. Kinetics of Exchange Between Americium (III) and Europium Ethylenediaminetetraacetate. J. Inorg. Nucl. Chem. 1973, 35(12), 4255.

10. Williams K. R.; Choppin, G. R. J. Kinetics of exchange of trivalent actinide ions with europium ethylenediaminetetraacetate. Inorg. Nucl. Chem. 1974, 36, 1849.

11. Muscatello, A. C.; Choppin, G. R.; D'Olieslager, W. Kinetics of dissociation of trivalent actinide chelates of trimethylenediamine-N,N,N',N'-tetraacetic acid [TMDTA]. Inorg. Chem. 1989, 28(6), 993.

12. Glentworth, P.; Wiseall, B.; Wright, C. L.; Mahmood, A. J. A Kinetic Study of Isotopic Exchange Reactions Between Lanthanide Ions and Lanthanide Polyaminopolycarboxylic Acid Complex Ions. I. Isotopic Exchange Reactions of Ce(III) with Ce(HEDTA), Ce(EDTA)-, Ce(DCTA)-, and Ce(DTPA)2-. J. Inorg. Nucl. Chem. 1968, 30(4), 967.

51

13. Laurenczy, G.; Brücher, E. Aminopolycarboxylates of Rare Earths. 12. Determination of Formation Rate Constants of Lanthanide(III)--tetraacetate Complexes via a Kinetic Study of the Metal-Exchange Reactions. Inogr. Chimi. Acta, 1984, 95, 5.

14. Southwood-Jones, R. V.; Earl, W. L.; Newman, K. E.; Merbach, A. E. Oxygen-17 NMR and EPR studies of water exchange from the first coordination sphere of gadolinium(III) aquoion and gadolinium(III) propylenediaminetetraacetate. J. Chem. Phys., 1980, 73 (12), 5909.

15. Cossy, C.; Barnes, A.; Enderby, J. The hydration of Dy3+ and Yb3+ in aqueous solution: A neutron scattering first order difference study. Journal of Chemical Physics 1989, 90 (6), 3254.

16. Choppin, G. Complexation kinetics of f-element and polydentate ligands. Journal of Alloys and Compounds 1995, 225, 242.

52

Chapter 4

Trans-Lanthanide Study Discussion

The previous chapter showed the results of kinetic measurements for the complexation reaction between the lanthanides and three PAPC ligands DTPA, EDTA, and HEDTA in 1 M

NaLac at three different acidities and 2 M total ionic strength. The similarities in the structures and binding groups of these ligands call for further comparison between the results. DTPA is octadentate with five carboxylate groups and three amine binding sites, EDTA is hexadentate with 4 carboxylate and two amine binding sites, and HEDTA is pentadentate with two amines and three carboxylate groups with the 2-hydroxyethyl group a potential weak sixth donor. In moving from DTPA to EDTA, there is a reduction of two coordination sites (one amine, one carboxylate) and moving from EDTA to HEDTA results in the reduction of one carboxylate binding site. These systematic changes to the coordination environments should lead to results from the kinetic experiments that lend themselves to interpretation of the effects of the different binding groups and the changing size cation of the coordination pocket.

Trans-lanthanide Trends

Figure 4.1 shows the trans-lanthanide trend for the complex formation rate constant with all three ligands investigated. It is notable that despite significantly different rates between the ligands and significant differences in thermodynamic stability of their metal complexes, each displays a similar trend in the behavior across the lanthanide series. An initial increase in the rate is seen at the middle of the lanthanide series followed by a fairly steady decrease in rate going through the heaviest lanthanides. The size of the initial increase and subsequent decrease is different for each ligand however. In the 1.0 M NaLac and 0.9 M HClO4 experiments the DTPA

53 reactions had the fastest rates overall but the increase in rate from the lightest lanthanide investigated, Pr3+, to the fastest metal in that series of experiments, Eu3+, was only a factor of about 1.5. The EDTA system exhibits an increase in rate of a factor of about 2 over the same range of metals. This suggests that while the additional binding sites of DTPA allow for faster overall reaction, the smaller binding pocket of EDTA is better able to take advantage of the changing coordination environment going across the series. In the HEDTA system the rates stay fairly constant from Pr3+ through the mid-lanthanides. This difference between the EDTA and

HEDTA trans-lanthanide trends suggests that the loss of a single carboxylate binding group results in the coordination environment change at the mid-lanthanides being inconsequential for

12000

10000

8000

6000

)

-1

s -1

4000

(M f

k 2000

100 50 0 0.84 0.86 0.88 0.90 0.92 0.94 0.96 1/r (CN=9, Angstroms-1)

Figure 4.1: Rate of complex formation across the lanthanides with (■) DTPA, (●) EDTA, and () HEDTA in 1.0 M NaLac 0.9 M HClO4 2 M ionic strength at 25 °C.

54 the complexation reaction in the high lactate system. This implies that in a high lactate system a ligand must be at least hexadentate to see an effect of the decrease in coordination number at the mid-lanthanides, and that pentadentate ligands do not cause enough steric crowding for there to be a benefit to more readily losing a coordinated solvent molecule.

In comparing results of individual lanthanides between the different ligands a much more significant effect of the different binging pockets is seen. The complex formation reaction of

DTPA with the lanthanides Pr3+, Eu3+, and Lu3+ are factors of 5, 4, and 2 faster than with EDTA respectively, however the complex formation reaction of EDTA with the lanthanides Pr3+, Eu3+, and Lu3+ are factors of 10, 20, and 8 faster than with HEDTA. This suggests that the loss of one carboxylate coordination site from EDTA to HEDTA has a much more significant effect on the rate of complexation than gaining one carboxylate and one amine coordination sites from EDTA to DTPA. With two carboxylate groups in fairly close proximity to one another, one of the binding groups may begin to interact with a lanthanide ion; holding the cation in a position such that the other carboxylate will be more likely to bind to the metal thus forming a large chelate ring that stabilizes the complex. With EDTA, this chelate ring would increase the opportunity for the other two carboxylate groups and as the amine groups to bind. This forms additional chelate rings that lead to the final complex. In losing one carboxylate group and replacing it with a hydroxyl the symmetry of the EDTA molecule is lost. The loss of this symmetry likely has a significant statistical effect on the initial interactions of the ligand with the lanthanide ions, reducing the likelihood of the initial reactions leading to the formation of the final complex.

As has been stated previously, there is a steady decrease in the rate of complex formation from the mid-lanthanides to the end of the series for all of the ligands investigated. It was mentioned in the previous chapter that this is the opposite of the effect one might expect as the

55 increase in the charge density of the lanthanides going across the series would suggest an increased interaction between the positively charged lanthanide ions and the negatively charged binding groups of the ligands. However this is not the effect observed in these studies. Starting from the fastest reaction of the ligands through to the end of the lanthanide series there is a factor of 24, 10, and 4 decrease in the rate of complex formation for DTPA, EDTA and HEDTA respectively. In a previous stopped-flow investigation of lanthanide complexation kinetics with

DTPA and EDTA in 0.3 M total lactate1, trends similar to those seen in the results from this work (figure 4.2) were observed in the kf values from light to heavy lanthanides. As this decrease in rate is seen in all of the ligands investigated, it suggests that the lactate medium might be

35000

30000

25000 )

-1 20000

s

-1

(M

f 15000 k

10000

5000

0 0.82 0.84 0.86 0.88 0.90 0.92 0.94 0.96 0.98 1/r (CN=9, Angstroms-1)

FIGURE 4.2: Rate of complex formation of the lanthanides with (■) DTPA, and (●) EDTA in 0.3 M

Lactot pH 3.6 (reproduced from [1]); (□) DTPA and (○) EDTA in 1.0 M Lactot pH 3.0 at 25 °C.

56 important in establishing the observed trend. The complex formation rates in 0.3 M Lactot are similar between DTPA and EDTA for the mid to heavy lanthanides and both ligands are significantly faster across the series compared to those in 1.0 M Lactot. The drop in the complex formation rate with increased total lactate of the system suggests that in such systems the lactate ion and/or lactic acid play an important role in the complexation kinetics. However, the effect of lactate in these reactions is not the only factor, as there is a significant difference in the complex formation rates between the different PAPC ligands in the 1.0 M Lactot system.

Figure 4.3 shows the trans-lanthanide trend for the complex dissociation rate with all the ligands investigated. It is seen that for the light and mid-lanthanides the kd values follow an

25

20

15

)

-1

(s d

k 10

5

0 0.84 0.86 0.88 0.90 0.92 0.94 0.96 1/r (CN=9, Angstroms-1)

Figure 4.3: Rate of complex dissociation across the lanthanides with (■) DTPA, (●) EDTA, and () HEDTA in 1.0 M NaLac 0.9 M HClO4 2 M ionic strength at 25 °C.

57 unexpected trend between the ligands where DTPA > EDTA > HEDTA. It is not unreasonable to expect the HEDTA to have the fastest dissociation rates and DTPA to have the slowest as each has the least and most coordination sites respectively. It would seem likely that between these similar PAPC ligand systems the metal complex with the most coordination sites would be the most kinetically inert to dissociation and the complex with the fewest would be the most kinetically labile. It is unclear why the kd values for light lanthanides with the different PAPC ligands would display this trend, but the lactate media may be a contributing factor. Figure 4.4

1 shows the previously published kd values in 0.3 M Lactot at pH 3.6 in comparison to the kd

18

16

14

12

10

)

-1

(s 8

d k 6

4

2

0 0.82 0.84 0.86 0.88 0.90 0.92 0.94 0.96 0.98 1/r (CN=9, Angstroms-1)

FIGURE 4.4: Rate of complex formation of the lanthanides with (■) DTPA, and (●) EDTA in 0.3 M

Lactot pH 3.6; (□) DTPA and (○) EDTA in 1.0 M Lactot pH 3.0 at 25 °C.

58 values obtained from the 1.0 M Lactot pH 3.0 system. The same observation of faster dissociation rates for DTPA over EDTA for the lightest lanthanides is seen in the 0.3 M Lactot system as in the 1.0 M Lactot system, although to a lesser extent. Additionally, while the difference in the pH between the two systems must be considered, the kd values seem to be reaching a lower limit for the heaviest lanthanides. Again, care must be taken to not over interpret the results of the extrapolative technique used obtain these values.

Isokinetic Relationship

To further analyze the 1.0 M total lactate systems the results of the temperature studies of each ligand can be compared. In the previous chapter tables 3.3, 3.6, and 3.9 report the activation parameters for the complex formation reaction in 1.0 M NaLac/0.9 M HClO4 and 2 M ionic strength between the lanthanides Nd3+, Eu3+, Tb3+, Ho3+, Tm3+, and Lu3+ with DTPA, EDTA and

HEDTA respectively. It was noted that despite having very different rates of complex formation at 25 °C the activation parameters for a given ligand were very similar between the different lanthanides. This same condition is observed between the different ligands as well as different lanthanide ions. Table 4.1 compares the enthalpies and entropies of activation of Nd3+, Eu3+,

Tb3+, Ho3+, Tm3+, and Lu3+ between the three ligands.

Table 4.1 highlights the surprising similarity in the activation parameters for the different lanthanides and PAPC ligands. Comparing the fastest and slowest lanthanides in this study, Eu3+

3+ 3+ and Lu , shows the unexpected nature of these results. The kf value at 25 °C for Eu with

DTPA is almost two orders of magnitude greater than the value with HEDTA however the

59 activation enthalpies for those reactions are identical within experimental uncertainty. The same

3+ observation is made with the Lu results where the kf value with HEDTA is about ten times slower than with DTPA. As mentioned in the previous chapter, the similar activation energies within the results for a single ligand do not readily explain the difference in the rates measured, and in comparing the results between the ligands there is an even greater difference in the rates and yet the activation enthalpies are very close to one another, some being within error. As the

ΔH‡ values do not readily explain the difference observed in the rates of reaction, the entropies of

DTPA EDTA HEDTA

-1 -1 Nd kf 6920 (10) 1310 (33) 120 (5) M s (25 °C) ΔH‡ 95 (4) 99 (11) 87 (3) kJ/mol ΔS‡ 146 (13) 148 (35) 86 (10) J/mol*K

-1 -1 Eu kf 11300 (90) 2650 (10) 130 (1) M s (25 °C) ΔH‡ 96 (3) 92 (2) 99 (2) kJ/mol ΔS‡ 156 (10) 128 (7) 126 (4) J/mol*K

-1 -1 Tb kf 7370 (50) 2830 (50) 110 (1) M s (25 °C) ΔH‡ 97 (2) 95 (3) 93 (3) kJ/mol ΔS‡ 155 (8) 138 (10) 105 (9) J/mol*K

-1 -1 Ho kf 2870 (20) 1810 (20) 70 (1) M s (25 °C) ΔH‡ 88 (2) 97 (2) 91 (9) kJ/mol ΔS‡ 117 (5) 142 (7) 96 (30) J/mol*K

-1 -1 Tm kf 1420 (40) 840 (50) 55 (11) M s (25 °C) ΔH‡ 80 (6) 96 (4) 88 (5) kJ/mol ΔS‡ 85 (21) 134 (12) 85 (17) J/mol*K

-1 -1 Lu kf 470 (10) 280 (10) 37 (7) M s (25 °C) ΔH‡ 107 (4) 96 (4) 105 (4) kJ/mol ΔS‡ 164 (15) 126 (12) 135 (14) J/mol*K Table 4.1: Activation enthalpies, entropies, and rate constants at 25 °C for the complex formation reaction between the Ln’s and PAPC ligands in 1.0 M NaLac, 0.9 M HClO4 and 2 M ionic strength.

60 activation must be examined for a compensating effect on the reaction rates. However this analysis does not yield an immediate explanation as the ΔS‡ values for the previously mentioned reactions of Eu3+ and Lu3+ with DTPA and HEDTA are identical within experimental uncertainty.

115

110

105

100

95

90

(kJ/mol)

‡ H

 85

80

75

70 60 80 100 120 140 160 180 S‡ (J*mol-1*K-1)

Figure 4.5: Isokinetic relationship for the complexation reaction between the lanthanides Nd3+, Eu3+, Tb3+, Ho3+, Tm3+, and Lu3+ with (■) DTPA, (●) EDTA, and () HEDTA.

In analyzing the ΔH‡ and ΔS‡ values for compensating changes between the different reactions, it is more instructive to examine all of the available data congruently by looking at the dependency of ΔH‡ on ΔS‡ in an isokinetic plot (figure 4.5). It is seen in figure 4.5 that there is a linear relationship in the dependency of ΔH‡ on ΔS‡ for the forward complexation reaction with the lanthanides for a given PAPC ligand. In an isokinetic relationship the slope of the line will be

61 in terms of absolute temperature and at that temperature the reactions in the series will occur at the same rate, referred to as the isokinetic temperature.2 The isokinetic temperatures from figure

4.5 are 266 ± 52 K for DTPA, 255 ± 107 K for EDTA and 315 ± 33 K for HEDTA. Care must be taken in the interpretation of these plots as the sizeable errors associated with it are a result of the compounding effect from the significant number of experiments and calculations required to obtain these results. The presence of an isokinetic relationship suggests that the reactions in a particular series are occurring via a similar mechanism.3 The somewhat similar isokinetic temperatures and similar activation parameters between the reactions of the lanthanides and the different PAPC ligands suggest that not only do the lanthanides complex with a given PAPC ligand via a similar mechanism, but that each of the PAPC ligands investigated complex with the lanthanides through a similar reaction mechanism as well. Given the structural similarities between the ligands, this conclusion is perhaps not remarkable.

Of the ligands investigated EDTA and HEDTA would seem the most likely to interact with metal ions via a similar mechanism, as the molecules are identical save for one of the chelating acetate groups on EDTA being replaced by a hydroxyl group in HEDTA. DTPA is a much larger chelate which takes up more space around the lanthanide ions and it has more points that can interact with a metal ion. Often in complexation reactions of multidentate ligands, the formation of the chelate ring is the rate determining step.4 It is seen in these results that each

PAPC ligand complexes with a lanthanide ion at significantly different rates. However the activation parameters suggest a similar mechanism for all the complexation reactions in these systems. Of these different reactions the only consistent thing between them is the high concentration of total lactate in the system. This suggests that some form of lactate plays a significant role in these reaction kinetics, either the lactate ion, lactic acid or both.

62

Unfortunately the results presented to this point do not provide a great deal of insight on the transition state of these complexation reactions. However the rate constant results can be further analyzed to provide a better understanding of these reaction systems through the calculation of a “kinetic equilibrium constant.”

Calculated Stability Constants

Under the restriction of microscopic reversibility, in which the forward and reverse reactions must occur via the same reaction pathway, a condition-specific “equilibrium constant” can be calculated as a ratio of kf to kd. Table 4.2 lists the logarithm of this ratio for all the lanthanides with DTPA, EDTA, and HEDTA in the 0.9 M HClO4 system. These values can then be compared to equilibrium constants calculated for the complexation reactions using the available thermodynamic stability constant data from the literature.5

Assuming that the most prominent protonated species of the ligands at the average pH of this study are the reactants of importance, the chemical equilibria for the complexation of the hydrated lanthanides with DTPA, EDTA and HEDTA are as follows

(4.1)

(4.2)

(4.3)

The respective equilibrium constants for these systems are

(4.4)

(4.5)

(4.6)

63

Using the DTPA case as an example it can be seen that these equilibrium constants can be calculated using the known stability constants5 for the following reactions

(βLnDTPA) (4.7)

(βH3DTPA) (4.8)

β (4.9) β

These calculated equilibrium constants for DTPA, EDTA, and HEDTA are listed in table 4.2 and it can be seen that the values calculated from literature stability constants do not match with the kf/kd values. It should be noted that the equilibrium constants calculated from literature were done with stability constants obtained at 0.1 M ionic strength. This was done in order to have an internally consistent set of stability constants as data for all of the reactions at 2 M ionic strength were not available. However it is unlikely that the stability constants under 2 M ionic strength would differ enough from the 0.1 M data to account for the large discrepancy between the calculated equilibrium constants and the kf/kd ratios. The differences between these values are an indication that these systems are not adequately described by only the equilibrium expressions for deprotonation of the ligand and complexation between the ligand and lanthanide ion.

Considering the high concentration of lactate present in all of the systems investigated, it is not unreasonable to postulate a mechanism in which the lanthanide ions are released by AAIII then rapidly form the tris lactate complexes, which should be the dominant lanthanide lactate complexes in 1 M lactate media at pH 3.6. Assuming that Ln(Lac)3 is the dominant species, the correct equilibrium expressions become

(4.10)

64

(4.11)

(4.12)

(4.13)

(4.15)

(4.16)

Using DTPA again as an example it can be seen that these equilibrium constants can be calculated using the known stability constants for the following reactions

(βLnDTPA) (4.17)

(βH3DTPA) (4.18)

(βLnLac3) (4.19)

(KHLac) (4.20)

β (4.21) β β

To allow for determination of a non-condition-specific value for the EDTA and HEDTA systems, it is assumed that third lactate ion becomes part of the bulk solvent after dissociation and is not included in the calculation. Values of the calculated equlibrium constants that include the Ln(Lac)3 species can also be found in table 4.2. It is seen that these values are in much better agreement with the kf/kd ratios. This suggests that while the lanthanide lactate complexes do not compete with the lanthanide PAPC complexes thermodynamically, the lactate complexes appear to play a significant role in the kinetics of these systems.

65

Lactate Complexation

The analysis of the calculated kinetic and thermodynamic equilibrium constants suggests that the complexes that lactate forms with the lanthanides are important to the kinetics. The thermodynamic equilibrium constants are calculated assuming a tris-lactate lanthanide species, however depending on the concentration of lactate ion in solution different lanthanide lactate

2+ + species can form. The Ln(Lac)1 , Ln(Lac)2 , and Ln(Lac)3 species will be in a rapid equilibrium

+ with each other with the Ln(Lac)2 and Ln(Lac)3 species being most prevalent. The average number of lactate ions coordinated to a lanthanide can be described as:

(4.22)

Using the stability constants for the formation of the separate lanthanide lactate species eq. 4.22 can be expressed in terms of solely the concentration of lactate ion:

(4.23)

(4.24)

(4.25)

Applying the approximation that the pH was the same for every experiment across the lanthanide series with each ligand, the acidities of 0.9, 0.85, and 0.8 M HClO4 have pH values of 2.6, 2.8, and 3.0 respectively. Using these pH values to calculate the lactate ion concentration, eq. 4.25 allows for calculation of the average number of lactates bound to a lanthanide under the experimental conditions before mixing with the PAPC ligand (table 4.3).

66

In nearly all the complexation reactions between the lanthanides and the ligands DTPA and EDTA, a decrease in the value of kf was observed with a decrease in the total acidity of the system (Tables 3.1 and 3.4). The decrease in acidity can be interpreted as an increase in the concentration of lactate ion, which will increase the ligand number of the lanthanide lactate complexes. In figure 4.6 a strong correlation is seen between a decrease in kf and an increase in the average number of lactate ions coordinated with Eu3+ in the reactions with both DTPA and

EDTA. This correlation was first observed in the 0.3 M total lactate study as the average number of coordinated lactate ions increased across the lanthanide seires.1 The same trans-lanthanide trend is seen in this study. Figures 4.7 4.8, and 4.9 show the kf values for the complex formation reaction of the lanthanides with DTPA, EDTA, and HEDTA respectively in 1.0 M NaLac at pH

2.6 as a function of the Lac:Ln ratio. Again a correlation is observed between a decrease in kf and an increase in the Lac:Ln ratio. This trend across the lanthanide series helps to explain why the complexation kinetics for all the PAPC ligands studied slow for the heavier lanthanides rather than increase as would be expected. The increase in charge density leads to greater interaction with the lactate ions and thus a higher Lac:Ln ratio.

This observation coupled with the highly positive ΔS‡ values indicates that lactate ion dissociation from the lanthanide ion is involved in the transition state of the complexation reaction. However these results do not allow for definitive distinction of the possible effects of lactic acid and hydrogen ion on the complexation kinetics, the concentrations of which are coupled to that of the lactate ion.

67

log(kf/kd) log(Kcalc) log(kf/kd) log(Kcalc) log(kf/kd) log(Kcalc)

Ln DTPA KLnDTPA KLnL3DTPA EDTA KLnEDTA KLnL3EDTA HEDTA KLnHEDTA KLnL3HEDTA Pr 2.52 (0.02) -1.48 3.43 2.24 (0.01) 0.65 1.89 1.39 (0.03) -0.47 0.77 Nd 2.80 (0.01) -0.96 3.67 2.37 (0.02) 0.86 1.82 1.55 (0.02) -0.2 0.76 Sm 3.15 (0.05) -0.23 4.43 2.75 (0.01) 1.41 2.4 1.91 (0.01) 0.23 1.22 Eu 3.32 (0.02) -0.19 4.38 2.76 (0.01) 1.6 2.5 1.85 (0.01) 0.26 1.16 Gd 3.28 (0.01) -0.19 4.58 2.81 (0.03) 1.7 2.8 1.78 (0.01) 0.12 1.22 Tb 3.20 (0.01) 0.14 4.79 2.93 (0.02) 2.22 3.2 1.72 (0.01) 0.2 1.18 Dy 3.23 (0.04) 0.25 4.59 3.01 (0.05) 2.65 3.32 1.61 (0.01) 0.18 0.85

68 Ho 3.25 (0.03) 0.21 4.41 3.25 (0.01) 2.91 3.44 1.36 (0.01) 0.2 0.73 Er 2.81 (0.03) 0.17 3.98 3.36 (0.07) 3.24 3.38 1.42 (0.06) 0.3 0.44 Tm 2.72 (0.04) 0.15 3.73 3.38 (0.18) 3.67 3.58 1.36 (0.09) 0.48 0.39 Yb 2.53 (0.02) 0.06 3.49 3.30 (0.02) 3.84 3.6 1.60 (0.01) 0.75 0.51 Lu 2.16 (0.01) -0.12 3.11 3.06 (0.08) 4.09 3.65 1.55 (0.10) 0.85 0.41

Table 4.2: Calculated kinetic equilibrium constants from the rate constants obtained in1.0 M NaLac 0.9 M HClO 4 2 M ionic strength. Thermodynamic calculated equilibrium constants calculated from literature data at 0.1 M ionic strength and 25 °C.

68

pH Ln 2.6 2.8 3 Pr 2.57 2.71 2.80 Nd 2.63 2.73 2.81 Sm 2.54 2.66 2.75 Eu 2.54 2.66 2.75 Gd 2.49 2.62 2.72 Tb 2.48 2.60 2.70 Dy 2.59 2.70 2.78 Ho 2.63 2.73 2.80 Er 2.73 2.81 2.87 Tm 2.79 2.86 2.90 Yb 2.81 2.87 2.91 Lu 2.85 2.90 2.93 Table 4.3: Average number of lactate ions coordinated to Ln3+ in 1.0 M NaLac at a given pH.

12000 11500 11000 10500 10000

9500 )

-1 9000

s

-1

(M f

k 2700

2400

2100

1800 2.55 2.60 2.65 2.70 2.75 Lac:Eu

Figure 4.6: Rate constant of complex formation of Eu3+ with (■) DTPA and (●) EDTA as a function of the average number of lactate ions coordinated to Eu3+ prior to mixing.

69

12000

10000

8000

)

-1

s -1

6000

(M

f k 4000

2000

0 2.45 2.50 2.55 2.60 2.65 2.70 2.75 2.80 2.85 2.90 Lac:Ln

Figure 4.7: Rate constant of complex formation of the lanthanides with DTPA as a function of the average number of coordinated lactate ions. 1.0 M NaLac, pH 2.6, at 25 °C.

70

3000

2500

2000

)

-1

s -1

1500

(M

f k 1000

500

0 2.45 2.50 2.55 2.60 2.65 2.70 2.75 2.80 2.85 2.90 Lac:Ln

Figure 4.8: Rate constant of complex formation of the lanthanides with EDTA as a function of the average number of coordinated lactate ions. 1.0 M NaLac, pH 2.6, at 25 °C.

71

160

140

120

100

) -1

s 80

-1

(M f

k 60

40

20

0 2.45 2.50 2.55 2.60 2.65 2.70 2.75 2.80 2.85 2.90 Lac:Ln

Figure 4.9: Rate constant of complex formation of the lanthanides with HEDTA as a function of the average number of coordinated lactate ions. 1.0 M NaLac, pH 2.6, at 25 °C.

72

As mentioned previously, the dependency of the Lac:Ln ratio was first observed for the rate constants of complex formation with DTPA and EDTA across the lanthanide series in the

0.3 M total lactate results.1 Despite the significant difference in the rates between the two systems similar trends in the kf dependency on the Lac:Ln ratio are observed in the DTPA studies

(figure 4.10). The difference in the kf values of the EDTA systems is too significant to permit a graphical comparison. However, DTPA and EDTA have very similar complexation rates in the

0.3 M Lactot system with the mid to heavy lanthanides while in the 1.0 M Lactot system the

EDTA rates are much slower than the DTPA rates. This observation suggests that while lactate ion ligation plays a role in the complexation kinetics at lower concentrations of Lactot, in the high

35000

30000

25000

20000

)

-1

s

-1

15000

(M

f k

10000

5000

0 2.4 2.5 2.6 2.7 2.8 2.9 3.0 Lac:Ln

Figure 4.10: Rate constant of complex formation with ligand vs. average number of bound lactates in (■) 0.3 M total lactate pH 3.6 with the lanthanides La3+-Lu3+ [1] and (●) 1.0 M total lactate pH 3.0 with the lanthanides Pr3+-Lu3+.

73

Lactot system the change in the medium is significant enough to increase the importance of the lactate ion interaction with the lanthanide cations. The lactate ion interactions with the lanthanides may in fact be the dominating interactions in the complexation kinetics of a high lactate system, as the activation parameters for all the ligands studied were found to have very similar values. Additionally, the change in the medium and increased importance of the role of the lactate ion may alter how the different PAPC ligands interact with the lanthanides as evident in the significantly different rates observed between DTPA, EDTA, and HEDTA in 1.0 M total lactate.

With regard to the rate of dissociation, there does not appear to be a correlation between the kd values and the Lac:Ln ratio (figure 4.11). This observation alongside the pH dependency

8

7

6

5 )

-1 4

(s d k 3

2

1

0

2.45 2.50 2.55 2.60 2.65 2.70 2.75 2.80 2.85 2.90 Lac:Ln

Figure 4.11: Rate constant of complex dissociation of the lanthanides with EDTA as a function of the average number of coordinated lactate ions. 1.0 M NaLac, pH 2.6, at 25 °C.

74

of kd seen across the lanthanide series for DTPA and EDTA, and for part of the lanthanide series for HEDTA, indicates that any effect by the lactate ion on the dissociation kinetics is not observable by these experiments. However these results are by no means definitive to such a condition and again care must be taken in the analysis of the kd values.

High Ligand Concentration

It was seen in the previous chapter that a saturation effect on the observed rate constant at high concentrations of DTPA takes place that is not present for EDTA or HEDTA in the concentration range studied (figure 3.13). Further investigation in the high DTPA concentrations at all of the acidities used in the trans-lanthanide studies found that as acidity increased so did the value at which kobs demonstrated this saturation (figure 3.14). The value at which the observed rate constant levels off would be expected to be the same for all acidities if the saturation effect was due only to the concentration of ligand. This suggested an additional complication in this system that would not be found in a simple saturation system. The analysis and interpretation of all the results presented strongly suggest that the lanthanide-lactate complexes play an important role in the complexation kinetics of these systems. As such it is reasonable to speculate that the saturation effect being observed in the Eu-DTPA system may be a result of the slow step in the reaction becoming completely dependent on lactate ion exchange kinetics. This is supported by the observed decrease in the saturation value for kobs with decreasing acidity, as the lactate ion concentration increases as acidity is lowered. This conjecture could be tested with some additional experimentation. In figure 3.13 it is seen that the kobs value for EDTA at 30 mM is similar to that of DTPA at 6 mM which is the last point studied before the saturation effect begins to take place. It is possible that at higher EDTA concentrations a similar leveling off of kobs could occur. If such were the case then saturation systems of DTPA and EDTA at the same

75

lactate ion concentration could be tested to see if the same value of kobs is obtained. If that were the result then it would be suggestive that under those conditions the rate being measured is that of the lanthanide-lactate exchange.

However before such experiments could be performed it is prudent to have a better understanding of the effect the species of lactate have in the systems already investigated. As the concentrations of lactic acid, hydrogen ion, and lactate ion are all interdependent on one another there is no direct way to measure the effect any one of these species has on the complexation kinetics. In order to isolate the effect of any one of those species on the kinetics a series of experiments with varying total lactate concentration and a tight control over pH are required. The next chapter will discuss the experiments performed to identify the individual roles of the different lactate species.

76

References

1. Nash, K. L.; Brigham, D.; Shehee, T. C.; Martin, A. The kinetics of lanthanide complexation by EDTA and DTPA in lactate media. Dalton Trans. 2012, 41, 14547.

2. Leffler, J. E. The enthalpy-entropy relationship and its implications for organic chemistry. J. Org. Chem. 1955, 20, 1202.

3. James H. Espenson. Chemical Kinetics and Reaction Mechanisms. 2nd Ed. Chapter 7 pp. 164. Mcgraw-Hill Inc., New York, 1995.

4. Laurenczy, G.; Brücher, E. Aminopolycarboxylates of Rare Earths. 12. Determination of Formaiton Rate Constants of Lanthanide(III)-ethylenediamine-tetraacetate Complexes via a Kinetic Study of the Metal-Exchange Reaction. Inorg. Chimi. Acta, 1984, 95, 5.

5. Martell, A. E; Smith, R. M. NIST Critically Selected Stability Constants of Metal Complexes Database; Version 8.0, 2004.

77

Chapter 5

Resolving the Mechanistic Role of the Lactate Buffer

In kinetic experiments it is a necessity that all reactant and product species in the system are known in order to confidently interpret the results. To determine the effect that one species has on the kinetics of the reaction of interest, the concentrations of every other component in the system must remain constant while the concentration of the species of interest is varied. This requirement presents an inherent difficulty in kinetic systems that involve a weak acid such as the lactic acid present in the previous lanthanide-PAPC complexation studies. The concentration of lactic acid ([HLac]) cannot be changed without also changing the concentration of the lactate ion ([Lac-]) and the pH, and therefore the effect that any one of these species has on the kinetics cannot be directly measured. As was described in the experimental chapter, a way to circumvent this limitation is to design experiments using several different concentrations of total lactate

- ([Lactot]) and adjusting each system to either have the same pH or the same [Lac ]. Completing these experiments at several values of pH and [Lac-] and then comparing the results between the systems should allow for interpretation of the kinetic role for each species in the lactate/lactic acid buffer system. However, considering the charge neutral nature of HLac, the species is unlikely to have a strong interaction with the trivalent Eu ion. As such the results here will likely lend themselves to mechanistic interpretations of [Lac-] and the hydrogen ion concentration

([H+]). The following experiments will investigate the Eu3+ complexation with DTPA in 1.0, 0.9, and 0.8 M [Lactot]. Errors on reported values correspond to a 95% confidence interval.

The experiments presented here require careful control over the pH, with each reactant solution being carefully adjusted to the requisite mV reading on an electrode calibrated based on

78

a Gran titration. This higher degree of pH control allows for adjustments to the plotted results that were not performed in the previous studies. The total concentration of DTPA used in these experiments is comprised of species with different degrees of protonation: DTPA5-, HDTPA4-,

3- 2- - H2DTPA , H3DTPA , H4DTPA , and H5DTPA. Given that the total concentration of DTPA is many times greater than that of Eu3+, it is reasonable to assume that the most prevalent protonated species of DTPA is likely to be the species interacting with the metal ion. Using the precise pH values and the DTPA protonation constants determined at 2 M ionic strength by

Thakur et. al.1, which were reproduced in this lab2, the speciation of DTPA at each pH used in the following experiments was determined (figure 5.1). The lowest pH used in these experiments

2- is 2.72 and it can be seen in figure 5.1 that H3DTPA makes up more than 50% of the total

90

80

70

60 H3DTPA

tot 50

H5DTPA

40 % DTPA % 30 H4DTPA

20

10 H2DTPA

0 2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 pH

Figure 5.1: Speciation of DTPA in 2 M ionic strength.

79

2- DTPA under these conditions. In addition to being the predominant DTPA species, H3DTPA on average has all five of its carboxylate groups deprotonated, increasing the likelihood of

- interacting with the europium ion over the next most prevalent DTPA species H4DTPA .

Changing the [DTPA] that is plotted with the observed rate constants from the total

[DTPA] to [H3DTPA], which is about half the total DTPA at pH 2.72, has a significant effect on the resulting slope, and in turn the kf value. Additionally, at the highest pH employed in this

3- study, 3.45, it can be seen that the species H2DTPA has increased to about 10% of the total

3- DTPA, which is not an insignificant amount of the total. H2DTPA also has all five carboxylates deprotonated, and therefore has at least the same amount of driving force to interact with the

2- metal ion as H3DTPA . Although the amine protons still need to be displaced for the formation of the final complex, the prior results in high [Lactot] suggest against these steps being rate limiting. Therefore, it is assumed these two species interact with the metal ion approximately the same in these systems. Therefore the concentration of DTPA plotted with the observed rate

2- 3- constants is the sum of the H3DTPA and H2DTPA species.

Constant pH

To obtain a clearer picture of the role of the hydrogen ion in the complexation kinetics of

3+ Eu and DTPA in the high [Lactot] media, the reaction was repeated at several constant pH values under the [Lactot] conditions of 1.0, 0.9, and 0.8 M; in this sequence, the free lactate ion concentration also varies. Figure 5.2 shows the results of the reactions run at a pH of 2.72, the pH at which 90% of [Lactot] is protonated. It is seen in figure 5.2 that with less total lactate/free lactate in the system the Eu-DTPA complex is formed more quickly. In addition it is seen that the kd value is approximately the same for each [Lactot] condition. Figure 5.3 shows the results

80

90

80

70

60

) 50

-1

(s

obs 40 k 30

20

10

0 0.000 0.001 0.002 0.003 0.004 0.005 H DTPA (M) 3,2

Figure 5.2: Observed rate constants for the reaction between Eu3+ with DTPA under a constant pH of 2.72 in 1.0 (■), 0.9 (●), and 0.8 () M total lactate. for the Eu-DTPA complexation reaction in 1.0, 0.9, and 0.8 M [Lactot] at a pH of 3.07, the pH at which 80% of the lactate in the system is protonated. Again it is seen that in the systems with less total lactate the Eu-DTPA complex forms more quickly and that the dissociation rate tends toward independence from [Lac-]. However, it is interesting to note that, although the increase in pH from 2.72 to 3.07 has increased the amount of H3,2DTPA in the system, the complexation reactions are slower at pH 3.07 than pH 2.72. It is also seen in figure 5.3 that again the rate constant of dissociation is approximately the same between the different [Lactot] conditions.

3+ Figure 5.4 shows the results of the Eu reaction with DTPA in the three [Lactot] at a pH of 3.30.

At this pH 70% of the lactate in the system is protonated. Once more the trend of faster complex formation for systems with less total lactate is observed, as well as the rates being slower in the

81

lower pH system despite more H3,2DTPA being present. However, the difference in the kobs values between pH 3.07 and 3.30 is a considerably less than that between pH 2.72 and 3.07. It can also be seen in figure 5.4 that once again the dissociation rate constant is approximately the same between the three [Lactot], supporting the concept that the complex dissociation is independent of [Lac-] under these conditions.

90

80

70

60

50

)

-1

(s 40

obs k 30

20

10

0 0.000 0.001 0.002 0.003 0.004 0.005 H DTPA (M) 3,2

Figure 5.3: Observed rate constants for the reaction between Eu3+ with DTPA under a constant pH of 3.07 in 1.0 (■), 0.9 (●), and 0.8 () M total lactate.

82

90

80

70

60

50

)

-1

(s 40

obs k 30

20

10

0 0.000 0.001 0.002 0.003 0.004 0.005 H DTPA (M) 3,2

Figure 5.4: Observed rate constants for the reaction between Eu3+ with DTPA under a constant pH of 3.30 in 1.0 (■), 0.9 (●), and 0.8 () M total lactate.

pH 2.72 pH 3.07 pH 3.30 -1 -1 -1 -1 -1 -1 -1 -1 -1 Lac tot kf (M s ) kd (s ) kf (M s ) kd (s ) kf (M s ) kd (s ) 1.0 M 21130 (520) 5.66 (0.26) 11530 (110) 3.13 (0.32) 9090 (250) 2.00 (0.60) 0.9 M 24530 (910) 4.67 (1.25) 13540 (280) 3.30 (0.43) 10550 (160) 3.16 (0.50) 0.8 M 27630 (680) 5.90 (0.84) 14840 (740) 5.19 (1.11) 12830 (630) 3.43 (1.43) Table 5.1: Rate constants of complex formation (kf) and dissociation (kd) under constant pH with different total lactate conditions.

Table 5.1 lists the rate constants of complex formation and complex dissociation in each

of the total lactate systems for all the constant pH conditions. The kd values for the [Lactot]

systems under each constant pH are seen to be constant within experimental error. The

consistency of these observations suggests that the rate constant of dissociation for the Eu-DTPA

complex is most significantly affected by the pH in the high [Lactot] systems. It is also seen that

83

the kd values decrease with decreasing pH, which fits with the previously observed acid catalyzed dissociation pathways of polyaminopolycarboxylate ligand complexes with the lanthanides.3

For the complex formation reaction it can be seen in table 5.1 that the kf values are different for every [Lactot] condition and pH. As the kf values are constantly changing no conclusions can be drawn from these results as to the role [H+] on the complex formation reaction. However some qualitative observations can be made, such as the significant decrease in the kf values between the pH systems of 2.72 and 3.07. Between those systems the protonation of the total lactate decreases from 90% to 80%. This is a fairly significant change in the composition of the system, however the constant pH data does not allow for determination if the

- - drop in kf is due to the decrease in the [HLac] or the increase in [Lac ]. Constant [Lac ] experiments will help differentiate the effect of each species, though it is unlikely that [HLac] has a significant role in the complexation kinetics.

Constant Lactate Ion

Experiments were done under the condition of constant [Lac-] to gain an understanding of the role of the lactate ion in the complexation kinetics. [Lactot] solutions of 1.0, 0.9, and 0.8 M were prepared such that the pH of each system brought the [Lac-] to a common value, thus yielding solutions with constant [Lac-] and variable pH. The results for the reaction of Eu3+ with

DTPA under the constant [Lac-] of 0.1 M are shown in figure 5.5. It is seen that the data from each [Lactot] cluster together and that each system behaves similarly. The rate constants of complex formation are effectively constant within the accuracy of the measurement; the rate

84

constants of complex dissociation follow the anticipated pattern of decreasing values with increasing pH.

90

80

70

60

50

)

-1

(s 40

obs k 30

20

10

0 0.000 0.001 0.002 0.003 0.004 0.005 0.006 H DTPA (M) 3,2

Figure 5.5: Observed rate constants for the reaction between Eu3+ with DTPA under a constant [Lac-] of 0.1 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 2.72, 2.77, and 2.83 respectively.

Figure 5.6 shows the results of the Eu-DTPA complexation reaction for the three [Lactot] systems under the constant [Lac-] of 0.2 M. Here the observed rate constants fall in a narrow cluster of values, however they are not as closely grouped as was observed with 0.1 M lactate ion. It is seen in figure 5.6 that again the kf values are constant between the different [Lactot] systems, however the kd values do not follow the expected trend with pH. Instead the rate constant of complex dissociation is seen to increase with increasing pH. Figure 5.7 shows the

3+ results for the complexation rate between Eu and DTPA for the three [Lactot] systems each

85

adjusted to a [Lac-] of 0.3 M. Once more it is seen that the rate constant of complex formation is constant across the three [Lactot] conditions. It is also seen that, as with the 0.2 M lactate ion systems, the kd values increase with increasing pH.

90

80

70

60

50

)

-1

(s 40

obs k 30

20

10

0 0.000 0.001 0.002 0.003 0.004 0.005 0.006 H DTPA (M) 3,2

Figure 5.6: Observed rate constants for the reaction between Eu3+ with DTPA under a constant [Lac-] of 0.2 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 3.07, 3.13, and 3.19 respectively.

86

90

80

70

60

50

)

-1

(s 40

obs k 30

20

10

0 0.000 0.001 0.002 0.003 0.004 0.005 0.006 H DTPA (M) 3,2

Figure 5.7: Observed rate constants for the reaction between Eu3+ with DTPA under a constant [Lac-] of 0.3 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 3.30, 3.37, and 3.45 respectively.

0.1 M Lac- 0.2 M Lac- 0.3 M Lac- -1 -1 -1 -1 -1 -1 -1 -1 -1 Lac tot kf (M s ) kd (s ) kf (M s ) kd (s ) kf (M s ) kd (s ) 1.0 M 21130 (520) 5.66 (0.26) 11530 (110) 3.13 (0.32) 9090 (250) 1.93 (0.48) 0.9 M 23280 (530) 3.61 (0.64) 11470 (300) 5.38 (0.96) 8820 (540) 4.92 (1.40) 0.8 M 23560 (100) 2.22 (0.12) 11500 (410) 9.92 (1.66) 9510 (740) 6.63 (2.48) Table 5.2: Rate constants of complex formation (kf) and dissociation (kd) under constant lactate ion concentrations with - different total lactate conditions. Respective pH values at 1.0, 0.9, and 0.8 M Lactot; 0.1 M Lac : 2.72, 2.77, and 2.83; 0.2 M Lac-: 3.07, 3.13, and 3.19; 0.3 M Lac-: 3.30, 3.37, and 3.45

Table 5.2 lists the rate constants of complex formation and complex dissociation for each

of the total lactate systems for each of the constant lactate ion concentrations. It is seen here that

- when [Lac ] is held constant, the observed kf values are either constant within the experimental

uncertainty or very close to one another in all the [Lactot] conditions. This suggests that the rate

87

of formation for the Eu-DTPA complex is most significantly affected by the lactate ion in the

- -1 -1 high [Lactot] systems. In table 5.2 it is seen that for 0.1 M [Lac ] the value of kf is ~23000 M s

- -1 -1 - and in 0.2 M [Lac ] the value of kf is ~11500 M s . Doubling [Lac ] decreases the kf value by half, suggesting that the rate of complex formation for the reaction of Eu3+ with DTPA has in inverse first order dependence on [Lac-].

As was noted earlier, the expected trend of increasing values for the rate constant of complex dissociation with decreasing pH is only observed in the 0.1 M Lac- systems. At 0.2 and

- 0.3 M Lac the opposite of the expected trend is observed, increasing kd values with increasing

- pH. It should be noted that in both of these [Lac ] systems the kd values at higher pH (0.9 and 0.8

M Lactot conditions) have fairly large relative uncertainties and thus care should be taken not to over interpret these values. It can be seen in table 5.2 that in the 1.0 M Lactot systems the kd values behave as expected across the three lactate ion concentrations.

Role of Lactic Acid

With experiments completed for the three total lactate concentrations under both conditions of constant [Lac-] and constant pH, it is beneficial to compare the systems to one another. Figure 5.8 shows the conditions of constant pH at 3.07 and constant [Lac-] of 0.2 M side by side. Comparing the constant pH system with the constant [Lac-] system provides information on the role of [HLac] in the complexation kinetics of the Eu-DTPA complex.

88

70 70

60 60

50 50

40 40

)

-1

(s

30 30

obs k

20 20

10 10

0 0 0.000 0.001 0.002 0.003 0.004 0.005 0.000 0.001 0.002 0.003 0.004 0.005 H DTPA (M) H DTPA (M) 3,2 3,2

Figure 5.8: Observed rate constants for the reaction between Eu3+ with DTPA under the conditions of constant pH of 3.07 (left) and a constant [Lac-] of 0.2 M (right) in 1.0 (■), 0.9 (●), and 0.8 () M total lactate.

As was discussed above, in the constant pH systems the kd value remains constant in each

- of the [Lactot] conditions and in the constant [Lac ] systems the kf value remains constant in each of the [Lactot] conditions. For the purposes of this pseudo-first order reversible reaction the only terms required to describe the system are the rate constants of the forward (kf) and reverse (kd) reactions. As demonstrated in figure 5.8, conditions have been found that will hold either kf or kd

- constant. kf is found to remain constant under conditions of constant [Lac ] for all of the [Lactot] investigated, and between those different [Lactot] systems both the pH and [HLac] are changing.

Similarly kd is found to remain constant in each [Lactot] system under conditions of constant pH,

- and between each of those [Lactot] systems both [Lac ] and [HLac] are changing.

89

In this analysis it is found that both the rate constants of complex formation and complex dissociation will remain unchanged under conditions where [HLac] is changing. This observation provides support for the initial assumption that [HLac] does not influence either the forward or reverse reactions and therefore does not play a role in the complexation kinetics of Eu3+ and

DTPA in high lactate media. This result is not surprising as lactic acid is a charge neutral species and has little driving force to interact with either the positively charged lanthanide ion or the negatively charged species of DTPA. However this interpretation of the kinetic data provides evidence to support the non-interaction of lactic acid in the complexation kinetics and as such will not require a term in the rate law.

High [DTPA] Under Constant Lactate Ion Conditions

In the initial investigations of the complexation kinetics between Eu3+ and DTPA in 1 M

Lactot, it was found that the behavior of the system changes at high [DTPA]. It was shown in chapter 3 (Results, figure 3.14) that the observed rate constant of the Eu-DTPA complexation reaction breaks from linearity past the initial DTPA concentration range studied of 2-6 mM and begins to level off around 15 mM DTPA. In order to gain more information on this change in the complexation kinetics experiments were performed at the higher concentrations of DTPA in the

1.0, 0.9, and 0.8 M [Lactot] systems under the conditions of 0.1, 0.2, and 0.3 M constant lactate ion. Figure 5.9 shows the results of the Eu3+ complexation reaction with DTPA under the

- constant [Lac ] of 0.1 M in the three [Lactot] conditions for the full range of DTPA concentrations studied. It can be seen that while the observed rate constants start out grouped together at low

[DTPA], they begin to separate at the higher concentrations of the ligand. However, under these conditions the reactions have become considerably faster, with observed rate constants of greater than 120 s-1. Interactions at this rate have reaction half times of less than 6 ms and therefore are

90

difficult to reproduce with similar behavior between individual shots of the stopped-flow as with the slower reactions. A result of this limitation is the sizable errors associated with the highest concentrations of DTPA in these 0.1 M Lac- systems, and although separation between the different [Lactot] systems can be seen these results should not be over interpreted, as the observed rate constants are all still constant within error. Figure 5.10 shows the complexation kinetics of

3+ Eu and DTPA across the full range of DTPA concentrations at each of the [Lactot] systems under the condition of 0.2 M Lac-.

160

140

120

100 )

-1 80

(s obs k 60

40

20

0 0.000 0.005 0.010 0.015 0.020 0.025 0.030 H DTPA (M) 3,2

Figure 5.9: Observed rate constants for the reaction between Eu3+ with DTPA under a constant [Lac-] of 0.1 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 2.72, 2.77, and 2.83 respectively.

91

160

140

120

100 )

-1 80

(s obs k 60

40

20

0 0.000 0.005 0.010 0.015 0.020 0.025 0.030 H DTPA (M) 3,2

Figure 5.10: Observed rate constants for the reaction between Eu3+ with DTPA under a constant [Lac-] of 0.2 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 3.07, 3.13, and 3.19 respectively.

92

160

140

120

100 )

-1 80

(s obs k 60

40

20

0 0.000 0.005 0.010 0.015 0.020 0.025 0.030 H DTPA (M) 3,2

Figure 5.11: Observed rate constants for the reaction between Eu3+ with DTPA under a constant [Lac-] of 0.3 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with pH values of 3.30, 3.37, and 3.45 respectively.

Under these conditions the systems are slower and more reproducible at the highest concentrations of DTPA, resulting in much smaller errors. As such the separation of the kobs

- values at different [Lactot] systems at high [DTPA] seen in 0.1 M [Lac ] is more clearly observable in the 0.2 M [Lac-] condition. In comparison to figure 5.6 where the observed rate constants between the [Lactot] systems fall close together but are still clearly separate with kobs for the [Lactot] systems increasing in the order 1.0 M < 0.9 M < 0.8 M, it is seen in figure 5.10 that the kobs values cross and at the highest [DTPA] the values increase in the order 0.8 M <

0.9 M < 1.0 M. Figure 5.11 shows the results of the Eu-DTPA complexation reaction in the three

93

- - [Lactot] systems at 0.3 M Lac . As in the 0.2 M [Lac ] experiments, a comparison of figure 5.7 with figure 5.11 shows that at low [DTPA] the kobs values start out close but separate with kobs values for the [Lactot] systems increasing in the order 1.0 M < 0.9 M < 0.8 M and at higher

[DTPA] the values cross and then increase in the order 0.8 M < 0.9 M < 1.0 M.

It is seen in comparing the different constant [Lac-] systems that as [Lac-] increases the value at which the observed rate constant levels off decreases. In the 0.9 M Lactot experiments the observed rate constant is about 130 s-1, 100 s-1, and 80 s-1 for the [Lac-] of 0.1, 0.2, and 0.3 M respectively. This inverse dependence of the observed rate constant under saturation [DTPA] conditions is not the same straight forward first order inverse dependence that was seen for the second order rate constant of complex formation at lower concentrations of DTPA. The consistent trend of the kobs values increasing in the [Lactot] systems 0.8 M < 0.9 M < 1.0 M suggest an additional factor comes into importance for the Eu-DTPA complexation kinetics at high concentrations of DTPA. This trend is observed at each constant [Lac-], where the only difference between the [Lactot] conditions is a small difference in pH. Therefore, the increase in

+ kobs with decreasing pH suggests a direct dependence of the rate on [H ] at [DTPA] saturation conditions.

Data Fitting

3+ The investigations of Eu and DTPA complexation kinetics in media of high [Lactot] with pH or [Lac-] held constant yielded results that reveal a good deal about how the system operates. Under conditions of constant pH it was revealed that the greatest influence on the rate constant of complex dissociation is the pH of the system. Operating under constant [Lac-] it was observed that at low [DTPA] the rate constant of complex formation has an inverse first order

94

dependence on [Lac-]. From these two observations it appears that HLac does not play a direct role in the complexation kinetics, as might be expected. Under the same constant [Lac-] conditions the limiting rate reached at the high [DTPA] limit was found to have a similar inverse dependence on [Lac-], as well an apparent direct dependence on [H+]. Given these observations the following expression is proposed to describe the complexation kinetics in these high [Lactot] systems:

Values for the fitting parameters, a, b, and c were determined iteratively via minimization of the residual sum of squares using the SOLVER function in Microsoft EXCEL for all the data in each lactate ion concentration. At low [DTPA] fitting parameter a has the greatest influence on the system and as DTPA concentration increases to the point of the rate reaching independence of

[DTPA] fitting parameters b and c come into importance, allowing for acid dependent and independent pathways. Statistical error analysis of the fitting parameters is provided in Appendix

A. Figures 5.12, 5.13, and 5.14 show the experimental data and the calculated fits for the systems run at 0.1, 0.2, and 0.3 M lactate ion respectively.

95

160

140

120

100 )

-1 80

(s obs k 60

40

20

0 0.000 0.005 0.010 0.015 0.020 H DTPA (M) 3,2

Figure 5.12: Experimental data and calculated fits for the constant [Lac-] condition of 0.1 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with respective R2 values for the fits of 0.976, 0.999, and 0.978. -3 Fitting parameters – a: 3813; b: 2.29•10 ; c: 1.22

96

120

100

80 )

-1 60

(s

obs k 40

20

0 0.000 0.005 0.010 0.015 0.020 0.025 H DTPA (M) 3,2

Figure 5.13: Experimental data and calculated fits for the constant [Lac-] condition of 0.2 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with respective R2 values for the fits of 0.964, 0.994, and 0.938. -3 Fitting parameters – a: 4173; b: 4.57•10 ; c: 2.19

97

90

80

70

60

) 50

-1

(s

obs 40 k 30

20

10

0 0.000 0.005 0.010 0.015 0.020 0.025 0.030 H DTPA (M) 3,2

Figure 5.14: Experimental data and calculated fits for the constant [Lac-] condition of 0.3 M in 1.0 (■), 0.9 (●), and 0.8 () M total lactate with respective R2 values for the fits of 0.968, 0.998, and 0.967. -3 Fitting parameters – a: 4961; b: 4.32•10 ; c: 4.62

It is seen in the above figures that the calculated fits have fair agreement with the experimental results. However, in order to obtain the best agreement between calculated and observed values, it was necessary to determine a set of fitting parameters for each data set at the three lactate ion concentrations. Therefore, the proposed model does not completely describe the system. As was shown in the experimental chapter, the observed rate constant of a pseudo-first order reversible reaction is the sum of the forward and reverse rate constants, however the proposed model does not include a term for the reverse reaction. This omission is intentional, as the only information gained on the rate constant of complex dissociation was that it does not change given a constant pH, and increases with a decrease in pH. However this behavior was not

98

seen for every set of conditions in the constant lactate studies. Therefore, further experiments would be necessary to gain a more complete understanding of the complex dissociation kinetics.

Additionally, accounting for the dissociation rate would require including two more fitting parameters, one each for acid dependent and acid independent terms. There is not enough information in the cumulative data set to adequately describe these terms and the starting values used in the fitting calculations would have to be entirely based on conjecture. However, the intercept values are relatively small compared to the observed rate constants for the majority of experiments. As such the collected data are comprised mostly of the complex formation reaction.

Therefore a model designed on the complex formation reaction alone should be a decent approximation of the system, and provide a strong starting point for further developing the rate law as more information becomes available.

It is important to keep in mind that there are inherent limitations to the interpretation of the data from these high lactate systems. The high concentrations of total lactate are likely causing a significant change to the aqueous media. Activity coefficients for lactic acid and lactate ion are unknown for such high [Lactot] and high ionic strength. All the analysis presented was performed under the assumption that the activity coefficients of the species involved were unity, and as such the analytical concentrations were used in the workup of the data. Since the activity coefficients are unknown in the non-traditional aqueous media used in these experiments, it is not entirely certain if the correct concentration values are being used in the analysis. As this limitation will be present for any model attempting to describe the system, it is reasonable to expect some deviation between experimental and calculated values.

99

Conclusion

By running several Eu-DTPA complexation experiments at different [Lactot] and adjusting the systems to have either constant pH or constant [Lac-] the roles of the species in the lactate/lactic acid buffer system in the complexation kinetics have been determined. The constant pH studies reveal that the rate constant of complex dissociation is directly dependent on [H+], while the constant [Lac-] systems showed that the rate constant of complex formation has an inverse first order dependence on [Lac-]. These two results provide evidence that [HLac] does not directly affect the complexation kinetics. Investigations of high [DTPA] under the condition of constant [Lac-] showed that the rate of complex formation eventually becomes independent of

DTPA and in this region it is found that the rate is most significantly affected by the concentration of lactate ion.

The observation of the Eu-DTPA complexation becoming independent of [DTPA] has interesting implications when considering a TALSPEAK-like solvent extraction system. In

TALSPEAK the DTPA concentrations used are greater than that used in these experiments, and therefore in the range of rate independence from DTPA. This suggests that under TALSPEAK conditions the only way to influence the complexation kinetics is to change the concentration of lactate ion in the system. This may help to explain why TALSPEAK lanthanide extraction kinetics are slow under low [Lactot]. However phase transfer kinetics are generally considerably different from homogenous aqueous phase kinetics and therefore this possible connection between the lanthanide-DTPA complexation kinetics and the lanthanide extraction kinetics will need to be explored in greater detail.

100

References

1. Thakur, P.; Mathur, J. N.; Moore, R. C.; Choppin, G. R. Thermodynamics and dissociation constants of carboxylic acids at high ionic strength and temperature. Inorg. Chim. Acta. 2007, 360, 3671.

2. Grimes, T. Coordination Chemistry of f-Elements in the TALSPEAK Process, Ph.D. dissertation, Washington State University, 2011, Ch. 2, 40.

3. Choppin, G. Complexation kinetics of f-element and polydentate ligands. Journal of Alloys and Compounds 1995, 225, 242.

101

Conclusion

As was mentioned in the introductory chapter, the inspiration for the preceding studies was the apparent conflict between the accelerative effect of increased total lactate in

TALSPEAK phase transfer kinetics1,2 and the decelerative effect of increased total lactate on La-

DTPA complex dissociation.3 Broadly put, this investigation was motivated by the observations in solvent extraction literature correlating the use of high buffer concentrations to facilitate acceptable phase transfer kinetics in TALSPEAK. The investigations to resolve the difference in these observations involved studying complexation kinetics of the polyaminopolycarboxylate ligands DTPA, EDTA, and HEDTA across the lanthanide series in 1.0 M total lactate at several pH conditions. Similar trends were observed across the lanthanide series between each PAPC ligand for both the rate constants of complex formation and complex dissociation. Despite disparate values between the ligands for the rate constants of complex formation, very similar activation parameters were found between both different lanthanides and different PAPC ligands.

This observation suggested that the mechanisms for complex formation between the lanthanides and PAPC ligands in high lactate media are very similar to each other. The observation that kf decreases as the average number of lactate ions coordinated to the lanthanide ion prior to the reaction increases suggested that the lactate ion is involved in the mechanism of Ln-PAPC complex formation. This hypothesis was confirmed by a series of experiments investigating Eu3+ complexation with DTPA in high total lactate media under conditions that allowed for constant lactate ion concentrations with changing pH values and constant pH values with changing lactate ion concentrations. With these experiments it was determined that the rate of complex formation has an inverse first-order dependence on the concentration of lactate ion and that the rate of complex dissociation is mainly dependent on the pH of the system. The kd observation from

102

3 those experiments does not match those found in the previous study where La-DTPA kd values were found to decrease with increasing total lactate concentrations between 0.1 M and 0.3 M at a constant pH of 3.5. As the observations of Ln-DTPA coordination kinetics in high total lactate agree more with the observed trend of increased phase transfer kinetics in the TALSPEAK process from high concentrations of total lactate, the difference in the previous Ln-DTPA complexation kinetics study and those presented here seems to be a result of the rather significant change in the aqueous medium going from comparatively low 0.3 M to 1.0 M total lactate.

The observation that the complexation rate between Eu3+ and DTPA in media of high concentrations of total lactate becomes independent of DTPA at high concentrations of the ligand has interesting implications with respect to the phase transfer kinetics in TALSPEAK.

From the results obtained in the series of studies presented here a potential explanation for the accelerative effect of increased lactate concentrations on the extraction rate may be offered. At low concentrations of DTPA it was observed that the rate of complex formation between the ligand and Eu3+ is inverse first-order dependent on [Lac-]. At high concentrations of DTPA, where the rate becomes independent of [DTPA], the most significant effect on the complexation rate is obtained by changing [Lac-]. By increasing the lactate ion concentration a decrease in the rate of complexation is observed. In a TALSPEAK solvent extraction system the concentration of DTPA will be well within the complexation rate saturation range, and therefore the only way to significantly affect the aqueous complexation rate between the lanthanides and DTPA is by changing the lactate ion concentration. It was observed that pH only has a slight effect on the complexation rate under DTPA saturation conditions and mainly affects the dissociation rate of the complex. As the pH of a TALSPEAK extraction system will be controlled for the best

103

performance of the cation-exchanging extractant HDEHP, increasing the lactate ion will be achieved by increasing the total lactate in the system, which is seen to increase the rate of lanthanide extraction. This effect can be explained in terms of the aqueous phase complexation kinetics. Starting with a Eu-DTPA system at equilibrium, the metal will be complexed by the ligand and non-extractable. Only when the metal dissociates from the ligand will it be extractable, and only if it is not re-complexed by another DTPA molecule. It has been observed that increasing the lactate ion concentration of the system slows the rate at which DTPA complexes with the lanthanides, and in the case of extraction increasing the lactate ion increases the amount of time in which the metal ion is free from DTPA and able to be extracted to the organic phase.

In considering the aqueous/organic interface of a TALSPEAK system, the interface will not likely be a smooth surface but rather quite irregular with aqueous or organic regions extending into the opposite phase. Considering the ethyl-ammonium linkages along the backbone of the Ln-DTPA complex, it can be pictured that the complex will orient itself with the more non-polar backbone toward the organic phase. In this orientation, when the lanthanide ion is eventually released it will be done so away from the organic phase and be less likely to meet with an extractant molecule before being complexed by DTPA again. In media of high lactate ion concentrations the lanthanide ion will rapidly form species with a Lac:Ln of 2:1 or 3:1, which will slow the rate of re-complexation by DTPA. Additionally the lactate ions coordinated to the lanthanide will be much more readily exchanged with an extractant molecule than the polydentate DTPA.

The observed behavior of lanthanide complexation with DTPA in high concentrations of total lactate seems to explain the increased rate of extraction of lanthanides in a TALSPEAK

104

system. However it is important to note the significant differences in homogeneous aqueous phase complexation kinetics and the kinetics of phase transfer. As such it is beneficial to examine the previous literature on the phase transfer kinetics under TALSPEAK conditions.

The kinetics of solvent extraction is influenced not only by the various chemical reactions that occur in the system, but is also a function of the rates of diffusion of the species that control the chemistry of the extraction process. Additional complications arise from the fact that there are two bulk phases where chemical reactions may occur, along with a liquid-liquid interface, which separates the two immiscible phases, or a thin volume region near the interface where reactions are possible as well. The overall rate of the phase transfer could be controlled by chemical reactions that take place in these regions, by the rate of diffusion by the solute species through these regions, or by a combination of both.

In the case of diffusion controlled extraction kinetics a distinction must be made between diffusion in the bulk aqueous or organic phases and diffusion through the layers surrounding the interface. In practically applied solvent extraction processes the bulk phases are stirred efficiently and as such the transport of reactants from the bulk to the region near the interface can be considered instantaneous. Therefore diffusion in the bulk phases may be neglected. However even if the bulk phases are stirred vigorously, diffusional processes can still influence the solvent extraction kinetics. Diffusion in the interface region can be described by several models and theories of varying complexity such as two-film, boundary layer, surface renewal, and penetration theory.4 Extraction kinetics that are controlled by the rate of these processes are said to take place in the diffusion region.

105

If one or more of the chemical reactions in the solvent extraction process are sufficiently slow compared to the rate of diffusion, the phase transfer kinetics occur in what is known as a kinetic regime. Under these conditions the kinetics can be described entirely in terms of rates of chemical reactions. However it is possible for both the diffusional processes and the chemical reactions to have similar rates, and in this case the solvent extraction kinetics take place in a mixed diffusional-kinetic regime. This is the most complicated case, requiring complex mathematical workup to describe the system.5 Often experiments are designed to avoid the inclusion of the diffusion terms, for example by increasing the stir rate of the bulk phases to reduce the thickness of the diffusion films.6

Experimental techniques for solvent extraction kinetics generally fall into one of the categories of highly stirred vessels, constant interfacial-area-stirred cells, rotating diffusion cells, moving drops, and short-time contact methods.4 Each of these methods attempts, to varying degrees, to control the hydrodynamic conditions near the interface and the interfacial area.

Studies into the solvent extraction kinetics of TALSPEAK have mainly used constant interfacial- area-stirred cell techniques,1,7,8 though there has been some recent development in moving drops techniques employing microfluidics.9 Initial investigations to the kinetics of TALSPEAK extraction by Kolarik et. al. found extraction rates to be rapid when extracting from a non- complexing aqueous media, i.e. with no DTPA present. The addition of DTPA significantly slowed the rate of extraction, and thus it was concluded that the formation or dissociation of the

Ln-DTPA complexes could be considered as the rate determining step.1 Their work also found that the inclusion of lactic acid increased the rate of extraction and they postulated that a lanthanide-lactate complex was involved in the transition from the DTPA complex to the

106

extracted species, however their results did not permit interpretation of the extraction mechanism.

In experiments by Danesi and Vandegrift investigating the kinetics of phase transfer for

Eu3+ and Am3+ between aqueous chloride solutions and n-dodecane solutions of HDEHP, the extraction rate data were interpreted using two models, one for a system in the diffusion regime and one for the kinetic regime. Although both models adequately described the rate data, the best fit was obtained when the kinetics were assumed to be entirely controlled by the chemical reactions.7 This result, alongside the observation by Kolarik et. al. on the Ln-DTPA complexation reaction, suggests that TALSPEAK reaction kinetics take place in the kinetic regime either in the aqueous phase or on the aqueous side of the interface, assuming adequate stirring of both phases.

Further work by Danesi and Cianetti into the kinetics of TALSPEAK extraction examined Eu3+ extraction by HDEHP from an aqueous media of polyaminopolycarboxylate ligands (DTPA or HEDTA) and lactic acid at pH 3.0. Their interpretation of the Eu3+-DTPA-

Lac-HDEHP extraction rate data called for interfacial complexation reactions between HDEHP and free Eu3+ ion, the Eu-DTPA species, a mixed Eu-DTPA-Lac species, and the Eu(Lac)2+ species. Unfortunately this model does not adequately describe the aqueous conditions of a

TALSPEAK system. As was seen in chapter 4, in the TALSPEAK like aqueous condition of 1.0

M total lactate at a pH of 3 the average number of lactate ions coordinated to Eu3+ is 2.75 (table

4.3). Therefore a model of the kinetics of extraction in TALSPEAK would need to account for

+ 3+ 2+ the Eu(Lac)2 and Eu(Lac)3 species while the free Eu ion and Eu(Lac) species would be negligible.

107

Additionally recent studies employing spectroscopic, luminescence, and thermometric experiments showed no formation of Ln-DTPA-Lac mixed complexes, at least not to a level detectable by those techniques.10 Given these observations the species interacting with HDEHP

+ at the liquid-liquid interface are likely the Eu(Lac)2 , Eu(Lac)3, and Eu-DTPA complexes. As it is observed that the kinetics of HDEHP extraction from a Ln-DTPA system are slow1 this implies that it the improved extraction kinetics are due to interactions between the Eu-lactate complexes and HDEHP at the interface. This is further supported by the observations in the series of studies presented here that increased lactate concentration increases the ligand number of the Ln-lactate complex which in turn slows the formation rate of the Ln-DTPA complexes.

This then increases the amount of time that the lanthanide ions will have to interact with HDEHP molecules at the interface.

If the accelerating effect of lactate in the TALSPEAK process is indeed due to slowing the rate of re-formation complex between DTPA and the lanthanides, the results described in chapter 3 on lanthanide complexation kinetics with PAPC ligands present a potential way to improve performance of the extraction system. Switching the aqueous complexant from DTPA to a PAPC ligand that has slower rates of complex formation with the lanthanides could result in faster phase transfer kinetics. It was observed that in high lactate media the rates of complex formation across the lanthanide series with DTPA were between one and three orders of magnitude faster than those with HEDTA (figure 4.1). It was also found that the rates of complex dissociation across the lanthanide series for the two ligands are much closer to one another, especially for the mid and heavy lanthanides (figure 4.3). Additionally, both DTPA and HEDTA were found to have decreasing kf values with increasing Lac:Ln ligand numbers (figures 4.7 and

4.9 respectively), though to a lesser extent with HEDTA. These observations in high lactate

108

media seem to suggest that HEDTA could improve the kinetics of a TALSPEAK-like system, and such a system could potentially require less total lactate based off the lesser effect of Lac:Ln ligand number seen with HEDTA.

Recent efforts to simplify the TALSPEAK process have involved examining potential alternatives to both the aqueous hold-back reagent and the organic extractant.11 It was found that exchanging DTPA with HEDTA and HDEHP with its structural analog 2-ethyl(hexyl) phosphonic acid mono-2ethylhexyl ester (HEH[EHP]) resulted in an extraction system that reached equilibrium faster, required less carboxylic acid buffer, and has much better agreement with models based on available thermodynamic data than TALSPEAK.12 This process was given the acronym TALSQuEAK (Trivalent Actinide-Lanthanide Separation using Quicker Extractants and Aqueous Komplexes). In a study comparing the effects of 0.1 M lactate and 0.1 M citrate on the system it was found that varying the carboxylic acid has little impact on the separation process overall, however slightly faster extraction rates were observed in the lactate system at pH

4.5, and at this pH the lactate ion would comprise most of the total lactate.12 While it is possible that the explanation for the accelerating effect of lactate in TALSPEAK proposed above could also occur in the TALSQuEAK process, there are additional factors present in TALSQuEAK that need to be considered. UV-Vis spectrophotometric analysis of Nd3+ in TALSQuEAK indicated that a not insignificant amount of a ternary Nd-HEDTA-Lac species is present in the separations.12 This species could play a role in the kinetics of the reactions at the interface that is not present in TALSPEAK. Therefore the acceleration of phase transfer kinetics with lactate present may not result from the same effect in TALSQuEAK as TALSPEAK.

The stopped-flow method of investigating lanthanide complexation kinetics employed in these studies has proven to be quite effective at probing the non-traditional aqueous media of

109

high lactate concentrations and has provided useful insights into a very complex system.

Additional information on such systems could potentially be obtained if a similar saturation effect on observed rate constants as seen with DTPA is also present at high concentrations of other PAPC ligands. If such were the case it may be possible to compare results between the ligands to examine Ln-lactate exchange kinetics, of course this may be limited by the rate of interaction between the lanthanides and the indicator dye AAIII. Additionally, the techniques used in the presented studies could be applied to other systems with high concentrations of carboxylic acid buffers, such as acetic acid or chloroacetic acid to investigate potential substituent effects. Substances similar to lactic acid, such as glycolic acid and alpha- hydroxyisobutyric acid could probe for structural or steric effects.

The overall impact of this work is an increased understanding of both the aqueous phase complexation kinetics as well as the phase transfer kinetics of the TALSPEAK extraction system. The activation parameters from the Ln-PAPC ligand complexation temperature studies showed an isokinetic relationship between DTPA, EDTA, and HEDTA across all the lanthanides studied, suggesting a similar mechanism for all the systems. The constant lactate ion/pH experiments showed that lactic acid does not participate in the complexation reaction and that the pH mostly effects the dissociation reaction. The overall conclusion from this work is that in systems with high concentrations of total lactate, such as in TALSPEAK, the complexation kinetics are almost completely governed by the lactate ion.

110

References

1. Kolarik, Z.; Koch, G.; Kuhn, W. Acidic Organophosphorus Extractants-XVIII The Rate of Lanthanide(III) Extraction by Di(2-ethylhexyl) Phosphoric Acid From Complexing Media. J. Inorg. Nucl. Chem. 1974, 36, 905.

2. Nilsson, M.; Nash, K. L. Manuscript in preparation.

3. Nash, K. L.; Brigham, D.; Shehee, T. C.; Martin, A. The kinetics of lanthanide complexation by EDTA and DTPA in lactate media. Dalton Trans. 2012, 41, 14547.

4. Danesi, P. R. “Chapter 5: Solvent Extraction Kinetics” Solvent Extraction Principles and Practice, Second Editon, Revised and Expanded. Marcel Dekker Inc., 2004, New York.

5. Danesi, P. R.; Vandegrift, G. F.; Horwitz, E. P. Simulation of Interfacial Two-Step consecutive Reactions by Diffusion in the Mass-Transfer of Liquid-Liquid Extraction of Metal Cations. J. Phys. Chem. 1980, 84, 3582.

6. Danesi, P. R.; Chiarizia, R. “The Kinetics of Metal Solvent Extraction” CRC Critical Reviews in Analytical Chemistry, Vol. 10, Iss. 1, 1980, CRC Press Inc., Boca Raton.

7. Danesi, P. R.; Vandegrift, G. F. Kinetics and Mechanism of the Interfacial Mass Transfer of Eu3+ and Am3+ in the System Bis(2-ethylhexyl) Phosphate-n-Dodecane-NaCl-Water. J. Phys. Chem. 1981, 85, 3646.

8. Danesi, P. R.; Cianetti, C. Kinetics and Mechanism of the Interfacial Mass Transfer of Eu(III) in the System: Bis(2-ethylhexyl)phosphoric Acid, n-Dodecane-NaCl, Lactic Acid, Polyaminocarboxylic Acid, Water. Sep. Sci. and Tech. 1982, 17, 969.

9. Nichols, K. P.; Pompano, R. R.; Li, L.; Gelis, A. V.; Ismagilov, R. F. Toward Mechanistic Understanding of Nuclear Reprocessing Chemistries by Quantifying Lanthanide Solvent Extraction Kinetics via Microfluidics with Constant Interfacial Area and Rapid Mixing. J. Am. Chem. Soc. 2011, 133, 15721.

10. Leggett, C. J.; Liu, Guokui; Jensen, M. P. Do Aqueous Ternary Complexes Influence the TALSPEAK Process? Solv. Extrn. and Ion Exch. 2010, 28, 313.

11. Braley, J. C.; Grimes, T. S.; Nash, K. L. Alternatives to HDEHP and DTPA for Simplified TALSPEAK Separations. Ind. Eng. Chem. Res. 2012, 51, 629.

12. Braley, J. C.; Carter, J. C.; Sinkov, S. I.; Nash, K. L.; Lumetta, G. J. The role of carboxylic acids in TALSQuEAK separations. J. Coor. Chem. 2012, 65, 2862.

111

Appendix A

An error analysis of the fitting parameters used in figures 5.12, 5.13, and 5.14 was performed using the support-plane method as described by Duggleby.1 In this analysis the fitting parameters, optimized by the minimization of the residual sum of squares (RSS), are adjusted and re-fit to test the sensitivity of a given parameter on the RSS. A fitting parameter was adjusted, for example a 10% increase in parameter a, and held constant at the new value. The

RSS minimization calculation was repeated, varying the other parameters to fit them to the new value of a. This procedure was repeated to find upper and lower limits that result in a 20% increase in the RSS. Starting with the optimized values of the parameters in each iteration, upper and lower limits were determined for the fitting parameters a, b, and c. Table A.1 lists each set of optimized parameters determined for fitting the 0.1, 0.2, and 0.3 M constant lactate ion experiments (Chapter 5) along with their upper and lower limits that increase the RSS by 20%.

The values in table A.1 help to define the relative importance of the individual fitting parameters in each set of conditions. Parameter a is seen to have considerable influence over the fit as the upper and lower limits for a at each [Lac-] is within 10% of the optimized value. The fit is considerably less sensitive to parameters b and c, however it is interesting to note how the sensitivity of each parameter changes with the changing conditions. Parameter b correlates with the acid independent pathway in the rate law, and this is reflected by the acidity conditions in each set of experiments. In the highest acidity (0.1 M Lac-) a change in b of about 50% is required to reach the upper or lower limits. In the lowest acidity (0.3 M Lac-) a change of about

35% in needed to reach the limits. The sensitivity of b increases with decreasing acidity, suggesting an increasing importance in the acid independent pathway. Parameter c, which correlates to the acid dependent term of the rate law, is seen to be the least influential. This is

112

reflected in the 0.2 M Lac- data, as a value of zero for lower limit only results in a 13% increase in the RSS. This decrease in sensitivity is likely a result of the decreasing importance of the acid dependent pathway as the acidity of the system decreases.

As stated in chapter 5, the proposed model does not completely fit the data. This is reflected in the changing sensitivity of the fitting parameters with the changing system conditions. However, the changing sensitivity provides insight on the relative importance of the terms in each system and may help to improve the model as more data becomes available.

113

a b c Upper Lower Upper Lower Upper Lower Optimized Optimized Optimized Lac ion Limit Limit Limit Limit Limit Limit 0.1 M 3813 4130 3527 2.29E-03 3.52E-03 1.10E-03 1.22 1.95 0.50 0.2 M 4173 4653 3756 4.57E-03 6.74E-03 2.51E-03 2.19 5.14 0.00*

114 0.3 M 4961 5457 4514 4.32E-03 5.96E-03 2.81E-03 4.62 8.46 1.06 Table A.1: Upper and lower sensitivity limits of calculated fitting parameters resulting in a 20% increase in the residual sum of squares. *Zero value for this parameter results in a 13% increase in RSS.

114

References

1. Duggleby, R. G. Estimation of the Reliability of Parameters Obtained by Non-linear Regression. Eur. J. Biochem. 1980, 109, 93.

115