ENHANCED ELEMENTAL OF FLUORINE AND CHLORINE VIA CHEMICAL IONIZATION IN THE AFTERGLOW OF AN INDUCTIVELY COUPLED

A Dissertation submitted to the Faculty of the Graduate School of Arts and Sciences of Georgetown University in partial fulfillment of the requirements for the degree of Doctor of Philosophy in

By

Joseph E. Lesniewski, M.S.

Washington, D.C. February 5, 2021

Copyright 2021 by Joseph E. Lesniewski All Rights Reserved

ii

ENHANCED ELEMENTAL MASS SPECTROMETRY OF FLUORINE AND CHLORINE VIA CHEMICAL IONIZATION IN THE AFTERGLOW OF AN INDUCTIVELY COUPLED PLASMA

Joseph E. Lesniewski, M.S.

Thesis Advisor: Kaveh Jorabchi, Ph.D.

ABSTRACT

There is an increasing need for trace-level analysis of fluorinated and chlorinated compounds in complex samples due to their prevalence among pharmaceuticals and environmental and food contaminants. Importantly, elemental analysis offers major quantitation advantages in this area by alleviating the need for compound-specific standards. However, the current state-of- the-art elemental analysis method, inductively coupled plasma-mass spectrometry (ICP-MS), faces fundamental limitations in sensitivity for F and Cl detection because the high ionization potentials of these elements reduce formation efficiencies of Cl+ and F+.

In this dissertation, new ionization chemistries are reported to overcome the fundamental limitations of ICP-MS. In the new methods, solutions of analytes are introduced into an ICP as aerosols, producing common gas-phase Cl and F small molecules from analytes. These species are then ionized via -neutral reactions in the atmospheric-pressure plasma afterglow where plasma gasses cool significantly.

Detection of chlorine as Cl- is demonstrated in one method with sensitivities twice as large as that of ICP-MS, highlighting the fundamental advantage of this approach. Notably, introduction of sodium and methanol into the plasma along with organochlorines significantly enhances formation of Cl-. Fundamental investigations of these effects reveal an ionization mechanism where sodium introduction into the plasma leads to formation of a gas-phase NaCl intermediate

iii from organochlorine compounds. NaCl then reacts with a reagent ion derived from methanol

- - oxidation and sodium, [Na(HCO2)2] , to produce [NaClHCO2] in the ICP afterglow. This complex releases Cl- inside the mass spectrometer upon ion activation.

NaF is hypothesized to form from fluorinated compounds in a similar fashion. However,

- - ionization of this intermediate by [Na(HCO2)2] is not thermodynamically favorable to produce F

. We therefore explore new reaction chemistries for ionization of NaF and demonstrate a

+ + mechanism for formation of Na2F via highly-favorable Na adduction in the plasma afterglow.

Notably, this approach offers two orders of magnitude improvement in ion formation and detection efficiency compared to current ICP-MS/MS methods for F measurement, further underscoring the fundamental advantages of chemical ionization in the plasma afterglow for analysis of high- ionization-potential elements. Finally, matrix tolerance for F detection in this approach is demonstrated through quantitation of fluoride in infant formula.

iv

ACKNOWLEDGEMENTS The research and writing of this thesis would not have been possible without the support from numerous people, and I would like to acknowledge their contributions.

First, I would like to sincerely express my gratitude to my advisor, Dr. Kaveh Jorabchi – His contributions and the opportunities offered in his lab have been instrumental to this work; I have benefited tremendously from the invaluable mentorship, vigorous scientific discussions, and continued support through all stages of this work.

I would like to recognize our industrial collaborators at PerkinElmer. In particular Dr. Hamid Badiei and Dr. Dickson Cheung (Version 6.0) – their contributions go above and beyond the level of ordinary collaborations

I would like to extend my gratitude to my examining committee – Dr. Yuye Tong, Dr. Sarah Stoll, and Dr. Edward Van Keuren – for their time and effort in critically reviewing this dissertation.

I’ve had the privilege of working alongside a remarkable group of colleagues in the Jorabchi lab. In particular, I want to thank the current group members – Dr. Kunyu Zheng, Dr. William McMahon, Michael Dolan, Sam White, and Wanqing Li – and all past group members for their contributions throughout my Ph.D. studies.

With gratitude, Joseph E. Lesniewski

v

TABLE OF CONTENTS

CHAPTER I: INTRODUCTION ...... 1

1.1. Scope of Dissertation ...... 1

1.2. Nontargeted Analysis ...... 1

1.3. Quantitation without Compound-Specific Standards: Elemental Methods ...... 3

1.4. Elemental Quantification of Chlorine and Fluorine...... 4

1.4.1. Elemental Methods ...... 4

1.4.2. Nuclear Magnetic Resonance (NMR) ...... 6

1.4.3. Inductively Coupled Plasma (ICP)-MS ...... 6

1.4.3.1. ICP Generation and Properties...... 8

1.4.3.2. Sample Introduction ...... 9

1.4.3.3. Ion Sampling ...... 10

1.4.4. Challenges of Fluorine and Chlorine Detection using Conventional ICP-MS Methods...... 12

1.4.5. Ionization with Alternative Gas Plasmas ...... 14

1.4.6. Negative-Mode ICP-MS ...... 14

1.4.7. Thermal Ionization of Fluorine as BaF+ ...... 15

1.5. Plasma Assisted Reaction Chemical Ionization ...... 17

1.6. Dissertation Outline ...... 18

1.7. References ...... 20

CHAPTER II: ATMOSPHERIC PRESSURE PLASMA ASSISTED REACTION CHEMICAL IONIZATION FOR ANALYSIS OF CHLORINATED COMPOUNDS SEPARATED BY LIQUID ...... 42

2.1. Introduction ...... 42

vi

2.2. Experimental Details ...... 45

2.3. Results and Discussion ...... 48

2.3.1. Effect of Carrier Gas Flow Rate ...... 48

2.3.2. Effect of Ionization Reagent ...... 51

2.3.2.1. Independent Introduction of Ionization Reagent ...... 51

2.3.3. Ionization Mechanism ...... 56

2.3.4. Effect of Plasma RF Power ...... 60

2.3.5. Effect of Analyte Solvent Composition ...... 61

2.3.6. Analytical Figures of Merit ...... 63

2.4. Conclusions ...... 67

2.5. References ...... 69

CHAPTER III: MECHANISTIC INSIGHTS INTO CHLORIDE ION DETECTION FROM THE ATMOSPHERIC-PRESSURE AFTERGLOW OF AN ARGON INDUCTIVELY COUPLED PLASMA ...... 74

3.1. Introduction ...... 74

3.2. Experimental Details ...... 76

3.2.1. Plasma Assisted Reaction Chemical Ionization (PARCI) ...... 76

3.2.2. Mass Spectrometers ...... 79

3.2.3. Argon Gas Velocity and Temperature Measurements ...... 80

3.2.4. Computational Methods ...... 82

3.2.5. Sample Preparation and Data Analysis ...... 82

3.3. Results and Discussion ...... 83

3.3.1. Analyte Breakdown ...... 84

vii

3.3.2. Sampling Plasma Products into the Reaction Tube ...... 86

3.3.3. Effect of Chemical Reactions ...... 89

3.3.3.1. Background ...... 91

3.3.3.2. Relation of Background Ions to Cl- ...... 94

3.3.3.3. Effect of Exhaust Flow and Carrier Gas Flow Rates on Clustering ...... 101

3.3.3.4. Implications for Cl Ionization ...... 103

3.3.3.5. Mechanistic Considerations for Enhancement of Cl- Detection by First-Order Clusters ...... 108

3.3.3.6. Application to Fluorine Detection ...... 109

3.4. Conclusions ...... 110

3.5. References ...... 114

CHAPTER IV: HIGH-SENSITIVITY ELEMENTAL MASS SPECTROMETRY OF FLUORINE BY IONIZATION IN PLASMA AFTERGLOW ...... 119

4.1. Introduction ...... 119

4.2. Experimental Details ...... 120

4.2.1. Sample Introduction and PARCI ...... 122

4.2.2. Mass Spectrometer ...... 122

4.2.3. Sample Preparation and Data Analysis ...... 123

4.2.4. Ab Initio Calculations ...... 124

4.2.5. Ion Chromatography ...... 125

+ 4.3. Na2F as a New Reporter Ion for F ...... 125

4.4. Cool Plasma Operation ...... 130

4.5. Sensitivity ...... 132

viii

4.6. Signal-to-Background (S/B) Ratio ...... 132

4.7. Compound-Independent Response ...... 133

4.8. Quantitation in Matrix...... 136

4.9. References ...... 140

CHAPTER V: SUMMARY AND FUTURE DIRECTIONS ...... 145

5.1. Summary of Core Findings ...... 145

5.2. Opportunities for Expansion of this Work ...... 146

5.2.1. Negative-Mode Phosphorous Detection using PARCI ...... 147

5.2.2. Positive-Mode Phosphorous, Sulfur, and Chlorine Detection using PARCI...... 149

5.3. References ...... 154

APPENDIX: PERMISSION DOCUMENTS ...... 157

ix

LIST OF FIGURES

Figure 1.1. Schematic diagram of a typical ICP-MS instrument...... 7

Figure 2.1. Schematic diagram of the ICP-based PARCI with dual cyclonic spray chambers and atmospheric pressure ionization interface coupled to an atmospheric sampling mass spectrometer...... 46

Figure 2.2. Effect of total carrier gas flow rate on a) chlorine sensitivity and b) reaction tube end temperature...... 50

Figure 2.3. Effect of ionization reagent (200 µM sodium acetate) solution flow rate and solvent type on chlorine sensitivity...... 52

Figure 2.4. Correlation plots for major background ions and chlorine sensitivity obtained at various reagent solution flow rates and solvent types with experimental conditions of Figure 2.3...... 58

Figure 2.5. Effect of plasma RF power on chlorine sensitivity and compound-independent response factors...... 59

Figure 2.6. Effect of methanol content in analyte solvent on chlorine sensitivity and compound- independent response factors...... 62

Figure 2.7. a) Gradient chromatographic elution of 2.5 µM chloramphenicol injected onto a C18 column and detected by PARCI-MS and b) calibration curve for chlorine (0.1-5 µM) based on peak areas from the chromatograms...... 66

Figure 3.1. Schematic diagram of ICP-PARCI ...... 78

Figure 3.2. Ratio of Cl response factors from flow injections of NaCl and chloramphenicol as a function of carrier gas flow rate at various exhaust velocity settings...... 85

Figure 3.3. Effect of aerosol carrier gas flow rate and exhaust flow rate at 1100 W on a) temperature at reaction tube outlet, b) argon gas velocity flowing out of the reaction tube, c) sampling efficiency into the reaction tube calculated from the ratio of measured gas flow rate at the end of the reaction tube (Equation 3.1) to the total aerosol carrier gas flow rate, and d) Cl- sensitivity measured as the average peak area from triplicate injections of 12.5 μM chloramphenicol...... 88

Figure 3.4. Effect of declustering potential on background ions at optimized ion source operating conditions for best Cl- sensitivity (2.0 SLPM carrier flow rate, 5.8 m/s exhaust velocity, 1100 W RF power)...... 93

x

Figure 3.5. Correlation of background ion intensities with Cl- sensitivity indicated by r2 values for linear regressions...... 96

Figure 3.6. a) r2 values for linear regression of background ions’ normalized intensities (collected at low declustering nozzle-skimmer potential) with 35Cl- sensitivity (measured at high declustering potential) at carrier gas flow and exhaust flow rates identical to those in Figure 3.3...... 97

Figure 3.7. a-h) TIC-normalized and, i-p) absolute intensities of formate and nitrite ions and their clusters with sodium formate as a function of carrier gas flow rate and exhaust setting...... 98

Figure 3.8. Background spectrum collected using soft ion sampling conditions (declustering potential of 10 V) at carrier gas flow rate of 2.4 SLPM and exhaust flow setting of 3.8 m/s...... 100

Figure 3.9. Effect of aerosol carrier gas flow and exhaust flow rates on average cluster - - order (n) for a) [Nan(HCO2)n+1] and b) [Nan(HCO2)nNO2] ions calculated using Equations 3.2 and 3.3...... 102

Figure 3.10. Relative sensitivities (flow injection peak areas) of chlorine-containing ions detected in soft ion sampling conditions (nozzle-skimmer potential difference of 10 V) upon injection of chloramphenicol...... 104

Figure 3.11. a) MS spectrum of clusters generated by nano-ESI from an aqueous solution of - sodium chloride and sodium formate, b) MS/MS spectrum of [NaHCO2Cl] at collision energy of - - 12 eV showing Cl as a major fragment ion, c) MS/MS spectrum of [NaHCO2Cl] at collision energy of 20 eV, indicating competitive generation of formate and chloride at higher collision - energies, d) MS/MS spectrum of [Na2(HCO2)2Cl] at collision energy of 20 eV, showing loss of NaHCO2 as the major fragmentation pathway from the second order cluster, e) MS/MS spectrum - of [Na3(HCO2)3Cl] at collision energy of 20 eV, showing successive loss of NaHCO2 as the major fragmentation pathway from the third order cluster...... 107

Figure 3.12. Schematic diagram summarizing the proposed Cl- ionization mechanisms...... 113

Figure 4.1. Schematics of a) ICP-based PARCI-MS, and b) conventional ICP-MS/MS for detection of F...... 121

Figure 4.2. Background spectrum of PARCI-MS with 80 μl/min 10-mM aqueous NaOH mixed with 250 μL/min HPLC water prior to introduction into the nebulizer...... 127

Figure 4.3. Background spectrum of PARCI-MS using aqueous 10 mM sodium acetate for post- flow injection (post-column) sodium addition...... 128

xi

Figure 4.4. Flow injections of 50 µM NaF (first triplicate) and 50 µM p-F-phe (second triplicate) + lead to formation of Na2F in PARCI-MS...... 129

Figure 4.5. Effect of aerosol carrier gas flow rate in ICP-MS/MS (BaF+) and PARCI-MS + (Na2F )...... 131

Figure 4.6. Ten replicate flow injections of 10 µM NaF used for estimation of limit of detection...... 134

Figure 4.7. Effect of aerosol carrier gas flow rate on ratio of flow injection peak areas for 50 μM injections of sodium fluoride and p-F-phe in PARCI-MS...... 135

Figure 4.8. Chromatograms for blank and spiked infant formula samples as well as fluoride standards in water using a) IC-PARCI-MS and b) IC-conductivity detector...... 138

Figure 4.9. Calibration curve for IC-PARCI-MS using NaF standards in water with a linear least- squares fit...... 139

- Figure 5.1. (a) Typical flow injections of 1 µM Na2HPO4 in H2O monitoring PO3 (m/z 79)...... 148

Figure 5.2. (a) Triplicate flow injections of 1 µM glyphosate in 50% methanol:H2O followed by + triplicate injections of 1 µM Na2HPO4 in 50% methanol:H2O monitoring Na3HPO4 (m/z 165)...... 151

Figure 5.3. (a) Triplicate flow injections of 50 µM Na2SO4 and 50 µM methionine in 50% + methanol:H2O monitoring Na3SO4 (m/z 165)...... 152

Figure 5.4. (a) Triplicate flow injections of 25 µM chloramphenicol (50 µM in Cl) in H2O and 50 35 + µM NaCl in H2O monitoring Na2 Cl (m/z 81)...... 153

xii

LIST OF TABLES

Table 2.1. Effects of easily ionizable elements added to methanol on chlorine sensitivity ...... 55

Table 2.2. Chlorine ion intensities from flat-top peak flow injection of chloramphenicol in water...... 63

Table 2.3. Comparison of chlorine detection limits and sensitivities ...... 65

Table 3.1. Operating parameters for ICP-based PARCI ...... 79

Table 3.2. Reaction thermochemistries calculated at the ωB97xD/aug-cc-pVTZ level ...... 94

Table 4.1. Reaction thermochemistries (kJ/mol) calculated at the ωB97xD/aug-cc-pVTZ level...... 126

- Table 5.1. Operating parameters for PO3 experiments ...... 147

Table 5.2. Operating parameters for simultaneous monitoring of P, S, Cl, and F in positive-mode...... 150

Table 5.3. Preliminary estimates of sensitivity and detection limits for simultaneous detection of P, S, Cl and F analytical ions using PARCI ...... 154

Table 5.4. Sensitivities and detection limits for ICP-MS/MS optimized for detection of a single element ...... 154

xiii

Chapter I

Introduction

1.1. Scope of Dissertation

Elemental quantification has continued to be an active area of research since the invention of ICP-MS over thirty years ago [1–6]. However, expanding these capabilities to nonmetals has been challenging because these elements suffer from low positive-mode thermal ionization efficiencies and isobaric interferences [7], resulting in high detection limits. In this dissertation, we address the fundamental ionization limitations of existing methods by developing efficient post-plasma chemical ionization for elemental quantification of nonmetals. Accordingly, a brief summary of the significance of nonmetal quantification, particularly with respect to nontargeted analysis, a succinct background of existing elemental mass spectrometry (ICP-MS) methods, and the current state-of-the-art in nonmetal element ionization are provided in the following sections to give context to the developments described in the subsequent chapters.

1.2. Nontargeted Analysis

Nontargeted trace analysis, or the search for unknown unknowns at trace level, has gained increased attention due to the value of determining the complete chemical composition of samples.

This type of analysis has found applications in numerous fields ranging from forensics [8, 9] to food [10–13], pharmaceutical [14, 15], and environmental analysis [16–18]. Nontargeted analysis provides valuable information in two scenarios: (1) when we don’t know what is likely to be in a given sample and (2) when we know that a parent compound was likely to be in a sample, but that compound was subjected to complex chemical transformations with unknown final products.

Either of these two scenarios can be found in environmental or food analysis [10–13, 16–18]. An example of the first scenario is 3-monochloropropane diol (3-MCPD), a likely carcinogen [19] that

1 was discovered via nontargeted analysis [20], as a byproduct of industrial food processes when fats are heated in the presence of salt or acid [21]. Scenario two is the case in the early stages of pharmaceutical development, where drug candidates are subjected to metabolism [22–25].

Recent developments in high-resolution mass spectrometry have enabled rapid and facile identification of unknown compounds through use of exact mass measurements [26–32], which when coupled to chromatography has made it the de facto technique for nontargeted analysis. The mass spectrometric methods for nontargeted analyses, however, fall short when it comes to providing rapid quantitation of detected compounds. Commonly used molecular ion sources, such as electrospray ionization (ESI) and atmospheric pressure chemical ionization (APCI), convert molecules to gas phase ions with efficiencies that depend on the properties of the specific molecules (i.e. gas-phase acidity/basicity for APCI and solution-phase chemistry for ESI) [33].

Consequently, response factors, defined as the number of ions detected for a given amount of analyte, are analyte-specific in these types of molecular ionization sources. Compound-dependent response factors necessarily mean that pure standards of the analytes are needed as a prerequisite to quantitation using molecular mass spectrometry. This requirement for compound-specific standards has created a significant bottle-neck in quantitative nontargeted analysis workflows [23,

34]. In other words, compounds of interest in complex samples may be readily detected and identified by exact-mass measurements using molecular ionization methods [35–37], whereas it is difficult to determine their concentrations without standards. Often, synthesis and purification of standards is needed for quantitation, requiring major research efforts and causing a significant lag time in between when new compounds are discovered and when they can be quantified [23, 34].

Rapid nontargeted quantitative measurements are therefore needed to provide quantitative metrics and to enable quantitative hypothesis formulation and testing in nontargeted investigations.

2

1.3. Quantitation without Compound-Specific Standards: Elemental Methods

Elemental analysis provides a potential solution for nontargeted quantitation. This is accomplished by employing instrumental methods that are sensitive to atoms rather than molecules. Thus, only one standard is needed for each element in these methods and the analytical response depends only on the amount of the element present, not the structure of the compound analyzed. Once the elemental concentration of a compound in the sample is determined, the molecular concentration may be inferred by considering molecular formulas.

Since elemental methods do not differentiate between molecules, separation via chromatography is needed to obtain concentrations of analytes that share common elements in a sample. Notably, common solvents in liquid chromatography such as water, methanol, acetonitrile, etc. contain significantly larger quantities of H, C, N, and O than the analytes they carry. Therefore, we will not be able to perform quantitation using these elements due to very high backgrounds of these elements contributed by the solvents.

To maximize our ability to detect and quantify compounds, nonmetals that are highly prevalent in biomedical and environmental applications are considered in this dissertation. F, Cl,

P, and S are abundant in such analytes with P and S being the most abundant of them [38]. Notably, reasonably sensitive (low ppb range limits of detection) universal quantification methods have recently been developed using ICP-MS/MS for S and P [39]. Conspicuously missing from these methods however are fluorine and chlorine, due to the fundamental challenges associated with detection and quantitation of these elements.

Fluorine and chlorine are nonetheless important given their abundance in pharmaceuticals

(30% of the top selling drugs contain F [40]) as well as in environmental pollutants [41–48] and food contaminants [21, 49]. Organochlorines are found in water disinfection byproducts, the

3 majority of the persistent organic pollutants (POPs) regulated by the Stockholm Convention [41], and nearly half of the compounds monitored by the FDA as pesticide residues [42]. The overwhelming prevalence of polyfluoroalkyl substances (PFASs), which are commonly used as oil and water repellent coatings [50], food packaging [49, 51], in aqueous film forming foams

(AFFFs) for fighting fires [52, 53], and a number of other consumer and industrial applications, has led to their detection in water supplies [53] as well as in biological serum [54]. Each of these classes of compounds is subject to complex transformation pathways either through environmental degradation [55–57], biological metabolism [25], or waste-treatment processes [43–45] leading to numerous transformation products. Notably, fluorine and chlorine are frequently retained in metabolites of drugs [58, 59] and other transformation products. For fluorine this is due in part to the high strength of the carbon-fluorine bond [60]. In fact, fluorine is intentionally added to pharmaceuticals to make particular metabolic sites less reactive [40].

In summary, elemental methods provide quantitation using universal standards, but are fundamentally challenged for analysis of Cl and F prevalent in many biologically and environmentally important compounds. Accordingly, this dissertation focuses on developing elemental quantification of chlorine- and fluorine- containing molecules separated with liquid chromatography.

1.4. Elemental Quantification of Chlorine and Fluorine

1.4.1. Elemental Spectroscopy Methods

Extensive work has been done in the field of elemental analysis using optical spectroscopy methods. However, fluorine and chlorine measurements are fundamentally difficult in atomic emission based spectroscopy due to the high excitation energies of these elements relative to metals [61]. There are additional challenges stemming from the most intense emission bands for

4 both fluorine and chlorine lying in the vacuum ultra-violet region (<190 nm), requiring elimination of molecular oxygen and water from the beam path of the spectrometer [62–64]. Short wavelengths are also more likely to encounter spectral interferences from overlap with molecular species, further requiring the use of high-resolution spectrometers [61]. Despite these challenges, quantitation can be accomplished for chlorine using inductively coupled plasma (ICP)-atomic emission spectroscopy with argon purged optics [62–65]. However, limits of detection are high

(ppm range) and significant matrix effects exist for these elements, making analysis of real samples difficult. Fluorine sensitivity for these methods is exceptionally poor due to high excitation potential leading to poor excitation efficiency, which in turn produces low emission intensity [66,

67].

To address the fundamental challenge posed by the high excitation energies of fluorine and chlorine and to avoid the issues in measuring emission in the vacuum ultraviolet range, methods relying on transient molecular species formed in high temperature environments have been developed. These methods work by adding a metal M to a sample to form a M-F or M-Cl molecule in a high-temperature reactor for analysis. Examples include AlF, AlCl, MgF, MgCl, GaF, GaCl,

InCl, CaF, and SrF formed in graphite furnaces and detected by molecular absorption spectroscopy

[61, 68–75] and SrF, CaCl, and CaF detected via molecular emission in laser induced breakdown spectroscopy (LIBS) [76–78]. Matrix effects are still limiting, however, particularly when samples contain significant quantities of metals other than the additive metal or of anions other than fluorine due to competition effects in the formation of the transient molecules. In addition, the need for a graphite furnace or for deposition of a sample on a substrate in the case of LIBS, means that, at best, coupling to chromatography must be conducted offline, creating a practical challenge.

Despite these limitations, continuum source molecular absorption spectroscopy using GaF has

5 been used in conjunction with molecular mass-spectrometry for quantitation of fluorinated pharmaceutical metabolites [72, 73], further illustrating the potential and the need for technologies capable of quantitation of fluorine and chlorine without compound-specific standards.

1.4.2. Nuclear Magnetic Resonance (NMR)

Quantitative NMR has been proposed as a method for elemental quantitation of fluorinated and chlorinated compounds since it offers nondestructive analysis, structural information, and the ability to analyze simple mixtures without chromatography. Further, fluorine and chlorine both offer 100% total abundance of NMR active nuclei (fluorine is monoisotopic (19F), while both 35Cl and 37Cl are NMR active), making them inherently more sensitive than elements like carbon, where only 1% of nuclei are NMR active. Fluorine NMR has an additional advantage over chlorine NMR in that backgrounds in biological and environmental samples are very low due to the miniscule number of naturally occurring fluorine compounds [79, 80]. There are however, practical limitations such as long acquisition times (detection limits improve with longer acquisition times ranging up to 24 hours) [23, 79–89] and matrix effects that cause line-broadening [79]. This complicates analysis of mixtures since long acquisition times mean that coupling to chromatography can only be accomplished offline. Consequently, in order to obtain the best detection limits, NMR often requires complex sample preparation procedures to remove matrix

[89], which further limits applications to untargeted screening and quantification since the design of sample preparation requires knowledge of the sample to ensure that it is not inadvertently removed with the matrix.

1.4.3. Inductively Coupled Plasma (ICP)-MS

6

plasma cooling sampling secondary auxiliary gas cone skimmer skimmer gas argon plasma

quadrupole injector

detector

nebulizer induction coil vacuum pumps

spray

chamber 7

drain

Figure 1.1. Schematic diagram of a typical ICP-MS instrument. Analytes in solution are converted to aerosol by the nebulizer.

Larger droplets are filtered out by the spray chamber and the smallest droplets are carried through the injector into the plasma torch by

the nebulizer gas where droplets are vaporized and molecules are broken down to atoms. High temperatures in the ICP cause efficient

thermal ionization of these atoms for most elements. Ions from the flow of the ICP are sampled directly into the vacuum of the mass

spectrometer where differential pumping and a series of skimmers are used to maximize sampling of the ions from the flow.

The modern state of the art in elemental analysis is inductively coupled plasma mass spectrometry (ICP-MS), owing to its superior sensitivity, selectivity, and limits of detection for most elements [13]. However, ICP-MS suffers from both fundamental and practical limitations with respect to quantification of nonmetal elements, and in particular chlorine and fluorine. To clarify these fundamental challenges and to lay out the rational for the approach to address them, an overview of the fundamental processes contributing to ion detection in ICP-MS are discussed below.

1.4.3.1. ICP Generation and Properties

A schematic of the ICP-MS instrument is shown in Figure 1.1. Inductively coupled plasmas are sustained using a cylindrical load coil which supplies radio-frequency (RF) power to a gas flow

(usually argon) at the downstream end of an ICP torch (composed of three concentric tubes) [7].

The oscillating current through the induction coil creates a magnetic field which oscillates rapidly along the axis of the ICP torch, causing ions and electrons in the field to accelerate [90, 91]. The plasma is ignited by seeding this field with electrons from a tesla spark, which accelerate rapidly through the outer annulus of the torch generating a large number of collisions with the neutral gas.

These collisions cause the temperature to rise, and more electrons and cations are generated by a fraction of the collisions, which also begin to accelerate in the oscillating magnetic field. An equilibrium is quickly established between ions, electrons, and neutrals at very high temperatures and a plasma is sustained.

In the ICP torch shown in Figure 1.1, the outermost flow is used as a cooling gas to protect the quartz walls from the plasma. The middle tube supplies most of the gas that sustains the plasma.

The center tube is called the injector through which analytes are introduced into the plasma. The

8 sample injection into the center of the plasma creates an annular plasma shape where the outer annulus (also known as the induction zone where the RF power is primarily coupled) can reach temperatures of ~10,000K, while the central channel where analytes reside typically reaches 5000-

6000K [7].

The high temperatures available in the ICP are the basis for its use as a thermal ion source.

There is sufficient energy at these temperatures to fully breakdown molecules to atoms even with significant solvent loading [7]. In these conditions any given element M exists as an equilibrium of M+ ions and neutral M atoms, as described by Reaction 1.1:

M ⇋ M+ + e− (푅푒푎푐푡푖표푛 1.1)

The ionization efficiency for a given element can be estimated using the theoretically calculated equilibrium constant for the ionization reaction above [7]. Interestingly, a near-complete ionization efficiency for most elements is estimated in a typical ICP, in particular elements with ionization potentials less than 8 eV (about half that of Ar: 15.76 eV) are ionized with 90-100% efficiency [7].

Notably, this characteristic implies that one element may serve as a standard for quantitation of other elements [92, 93]. Frequently, a single metal standard is used to calibrate for one or more other elements with similar ionization efficiency, usually when simultaneous quantification of many elements is needed [93], or when the metal of interest is rare (expensive) or dangerous

(radioactive) [92].

1.4.3.2. Sample Introduction

Despite the inherent plasma robustness and near 100% ionization efficiency of many atoms in an ICP, there are several processes that reduce the total detection efficiency of atoms in a sample by ICP-MS. The first step in an analysis is introduction of the sample into the plasma. As

9 mentioned earlier, the gas flow from the injector carries the analytes through the central channel of the plasma in the form of aerosolized droplets. These droplets are typically produced by concentric glass nebulizers and pass through a spray chamber to filter out the droplets > ~10 µm

[94]. The filtration reduces solvent loading in the ICP and minimizes plasma cooling by solvent.

It has been shown that large or fast moving droplets can actually survive the plasma [95], indicating excessive plasma cooling, which can lead to matrix effects and compromised thermal ionization efficiency.

As a consequence of droplet size filtration, transmission of analytes through the sample introduction process is reduced [96, 97] and can result in significant losses of analyte molecules before they ever reach the plasma. High efficiency nebulizers [98], direct injection nebulizers [99,

100], total consumption sample introduction systems [39, 100, 101], and different spray chamber designs [100, 102] have been implemented with the aim of bringing sample transmission efficiency closer to 100% by making smaller droplets and getting more of them into the plasma. Total consumption nebulizers and direct injection nebulizers are in fact able to achieve 100% sample introduction efficiencies, but at the cost of significantly limited solution flow rates in both cases

[99, 101].

1.4.3.3. Ion Sampling

The second process contributing to losses in atom detection efficiency is sampling of ions from the plasma into the mass spectrometer [7]. A diagram of the “three aperture” ion extraction interface for ICP-MS employed by modern instruments [103] is shown in Figure 1.1. Gas from the plasma is sampled directly into vacuum through a ~1.0-mm orifice in a sampling cone. The cone shape is designed to minimize sheath and boundary layer formation upon sampling the flow from

10 the plasma into vacuum [7]. Notably, the large pressure difference between the two sides of the sampling orifice creates a choked flow, where the amount of gas sampled from the plasma is determined by the geometry of the orifice and the pressure outside the sampling orifice. In these conditions, the amount of gas sampled is estimated to be 8.6 x 1020 atoms/s or roughly 1.9 SLPM

[7]. The aerosol gas sampling efficiency, defined as the fraction of the carrier gas that is sampled into the mass spectrometer, is therefore close to 100% because carrier gas flow rates of ~1.0 SLPM are typically used in ICP-MS. Deviation from 100% occurs due to efficient diffusion of analytes and central channel gas into the outer gasses at the high temperatures observed in the ICP.

Once the plasma gasses enter the first vacuum stage of the mass spectrometer, the gases expand and cool to temperatures of around 150 K, converting random thermal movement into a supersonic directed flow that expands into the vacuum. The free expansion into vacuum creates a region of low density (known as the zone of silence) forming a directed jet as well as shock waves.

A skimmer positioned inside the zone of silence samples the gas jet prior to shockwaves that cause the ions to scatter [7, 103]. This process represents a significant loss mechanism in ICP-MS since approximately 1% of gas that was sampled into the mass spectrometer is transmitted through the first skimmer [7]. A second skimmer is sometimes used to further reduce shockwaves, improve the ion/gas ratio, and limit the current of the ion beam for improved transmission through ion optics.[103].

Space charge effects in the directed jet region of the interface (downstream of the first skimmer) constitute yet another ion loss mechanism. It is of note that a large current of Ar+ (~1.5 mA) enters the vacuum along with analytical ions (such as Cl+ or F+). Space-charge effects arise when densely packed charges repel each other, defocusing the ion beam. Space-charge effect is

11 especially significant for ions with masses less than that of argon (such as F+ and Cl+), resulting in extensive defocusing of these ions and reduced transmission efficiency through the MS.

1.4.4. Challenges of Fluorine and Chlorine Detection using Conventional ICP-MS Methods

We now consider the challenges specific to chlorine and fluorine in ICP-MS, based on the thermal ionization process and ion loss mechanisms (sample introduction and ion sampling) discussed above. As mentioned above, space-charge effects are particularly detrimental to analysis of Cl+ and F+. A more impactful fundamental limitation in the analysis of these elements, however, is due to their high ionization potentials. The ionization potentials of chlorine and fluorine are

13eV and 17.4eV [104], respectively, much greater than the 8 eV ionization threshold discussed above. This results in very low ionization efficiency for Cl+ and F+ (0.9% and 0.0009% respectively) [7].

+ + Further adding to this challenge are abundant isobaric interferences such as H3O for F ,

34SH+ and 16O18OH+ for 35Cl+, and 36ArH+ and 36SH+ for 37Cl+. The complexity of the ICP interface to enable sampling of a 5000 K plasma with minimal distortion of plasma ion population means that instruments must be specifically built for this purpose. Therefore, a dedicated instrument is needed for elemental analyses, increasing the barrier for the widespread adoption of

ICP-MS in untargeted methods. More importantly, the complex interface has limited the progress in high-resolution and MS/MS in elemental mass spectrometry for resolving isobaric interferences, while advanced ion analysis techniques in molecular mass spectrometers are becoming more available as a result of recent advances in mass analyzer designs.

Despite these challenges, efforts utilizing conventional ICP-MS have been made to enable quantification of F and Cl. A portion of these efforts have focused on eliminating isobaric

12 interferences. F+ may be detected directly by ICP-MS if solvents (and therefore isobaric

+ interference form H3O ) are eliminated by replacing sample introduction with an electrothermal vaporizer [105]. However, this approach is of limited utility because of difficulties in online coupling to LC. Recently developed methods for chlorine quantification have employed collision

+ cells or triple quadrupole instruments with O2 or H2 reaction gasses to shift Cl to masses free of isobaric interferences, leading to improved limits of detection in the ppb range [106–113].

Additional methods have coupled sector-field mass spectrometers to ICP, resulting in ppb range detection limits for chlorine and making it possible to quantify fluorine using F+ with solvent in the plasma [114]. Even with the isobaric interferences resolved, detection limits for fluorine are still limited to the ppm range by poor ionization efficiency, indicating that the only potential advantage that HR-ICP-MS offers over optical methods is reduced matrix effects due to the high specificity of mass spectrometry.

Introduction of organic solvents (often used in liquid chromatography) into ICP raises additional challenges for quantitative detection of elements. Increased organic solvent loading cools the plasma, substantially reducing thermal ionization efficiency [115, 116], while increased transport efficiency in the sample introduction carries more analyte into the plasma [117, 118].

Further, chemical effects of carbon in the plasma produce enhancements in ionization efficiency through charge transfer or reduction of oxide formation by carbon species [39, 119, 120].

Accordingly, elemental response factors are altered with solvent composition, governed by the complex interplay of the aforementioned chemical and physical effects, consequently complicating quantitation using gradient chromatography where solvent composition changes during analysis.

There are two ways of dealing with the solvent effects that allow for quantitation of elements with

13 gradient liquid chromatography. One approach is to utilize a gradient-dependent calibration, as demonstrated for quantification of chlorinated pharmaceutical metabolites [106]. The other method is to buffer the plasma by adding carbon (e.g. methane) so that chemical and physical effects of carbon in the plasma don’t change significantly across the gradient [39].

Despite the efforts discussed above for F and Cl analysis in ICP-MS, the performance of conventional elemental MS instruments remains limited in terms of sensitivity due to the fundamental limitations of thermal ionization. Further improvement of these methods therefore requires a solution to the fundamental problem.

1.4.5. Ionization with Alternative Gas Plasmas

Early efforts to overcome the low thermal ionization efficiencies of F and Cl focused on a pure helium ICP [121], a pure helium microwave induced plasma [122], or a mixed gas helium- argon ICP [123] instead of a pure argon ICP. The basis of this is the higher ionization potential of helium (24.58 eV) relative to argon (15.76eV), meaning that a plasma sustained in helium will have higher electron temperatures (more energetic electrons), which could result in more extensive ionization of F+ and Cl+ [7]. However, helium plasmas had a number of drawbacks because of the nonthermal nature of these plasmas where the electron temperature is significantly higher than the gas temperature [7]. Low gas temperatures lead to enhanced formation of polyatomic interferences and matrix effects. Importantly, excessive quenching of the high-energy species in the plasma is observed upon solvent introduction [122, 124, 125], compromising the applicability of these plasmas for coupling to LC.

1.4.6. Negative-Mode ICP-MS

14

Given the inefficiency of thermal ionization in positive mode for F and Cl, negative mode ionization has been explored [126–128]. Negative mode ICP-MS experiments showed promising sensitivity and detection limits compared to positive mode. However, while these efforts overcame the fundamental ionization limits of positive mode ICP, they introduced a different fundamental limitation to ionization. Notably, ion kinetic energy measurements in negative mode ICP-MS showed that negative ions are not formed inside the plasma; rather, they are formed in the ion sampling interface during gas cooling via expansion into vacuum [126, 129]. In fact, this is not a surprising conclusion considering that the same thermal energy sufficient to ionize atoms to positive ions is enough to completely ionize F- to F atom and Cl- to Cl atom [7] since their electron affinities are only 3.4 eV [130] and 3.6 eV [131], respectively, compared to ionization potential of

5.1eV [104] for sodium which ionizes close to 100%.

Notably, the ionization reactions in low-pressure region of the MS where negative ions form are not efficient due to reduced collisions with electrons at low pressure. Moreover, the limits of detection in these efforts were compromised by excessive background from the isobaric interference 18OH- for fluorine, and very high chlorine backgrounds [126–128].

1.4.7. Thermal Ionization of Fluorine as BaF+

A recent effort to address both the fundamental (ionization) limitations and practical limitations (isobaric interferences) for fluorine analysis in ICP-MS is to convert fluorine to a positive ion that forms more favorably inside a hot plasma than F+. This is accomplished by introducing barium through the plasma concurrent with the F-containing analytes and carefully tuning plasma conditions for formation of BaF+ [132–136]. BaF+ is then resolved from Ba18OH+ using a triple quadrupole MS with oxygen as a reaction gas. Recently it was shown that this method

15 can be implemented on single quadrupole ICP-MS instruments as well, although with increased limits of detection due to high backgrounds from isobaric interferences [137].

Barium is specifically chosen over other metals due to its optimal combination of relatively high bond strength of M-F bond, low bond strength of M-O bond and low ionization potential of

M-F [133, 135]. These three parameters are critical for detection of M-F+ by ICP-MS since increased M-O bond strength results in production of more M-OH polyatomic species that interfere with fluorine detection, decreasing signal-to-background ratios and increasing limits of detection.

Increased M-F bond strength obviously improves the likelihood that the M-F polyatomic will survive in the hot plasma and lower ionization potential of M-F consequently increases the ionization efficiency for formation of M-F+. Experimental trends using various metals have supported the hypothesis that optimized analytical performance is achieved when M-O bond strength and M-F ionization energy are minimized, and M-F bond strength is maximized [135].

This approach creates a fundamental difficulty in the optimization landscape because of interplay between formation of BaF and BaO. Ba-F polyatomic formation is favored at lower plasma temperatures. However, such temperatures also increase formation of Ba-O and depletion of BaF, especially in oxygen-rich plasma environments that result from solvent introduction.

Accordingly, a very sharp sensitivity optimum is observed for BaF+ relative to typical elemental ions. BaF+ sensitivity drops by a factor of approximately 10 at carrier flow rates 0.1 L/min above or below the optimum [132, 134], compared with atomic ions like F+, Cl+, Br+, and I+ which lose only a factor of 2 in sensitivity for a change of 0.3 L/min above or below the optimum [114].

Higher carrier gas flow rates cool the central channel and decrease the residence time of analytes in the plasma [123]. Notably, the optimum flow for BaF+ detection is observed at a rate of 1.3

16

L/min, which is higher than the typical ~ 1 L/min used for analysis of metals in ICP-MS. Above this optimum value, the plasma becomes too cool to break Ba-O and Ba-F formation efficiency is lost, while lower flow rates prevent formation of Ba-F bond because of high temperatures.

Interestingly, an alternate explanation to losses at high carrier flow rates has been proposed based on reduction in Ba2+ population [134]. This hypothesis posits that the ionization mechanism for

BaF+ is an ion/ion reaction between Ba2+ and F-. This would also result in a sharp optimum since higher temperatures would neutralize F-, while cooler temperatures would reduce the ionization efficiency of Ba2+ on either side of the optimum.

Regardless of the mechanism, the steep optimization surface and the need for precise control of the plasma temperature detract from the robustness of the ionization method, leaving F elemental quantitation an unresolved analytical challenge. In fact, matrix effects have been reported for BaF+ method when other metals are introduced with the sample [132, 136].

1.5. Plasma Assisted Reaction Chemical Ionization

To overcome the fundamental challenges arising from thermal ionization of F and Cl, a new method relying on chemical ionization is developed in this dissertation. The new method is based on the general strategy of Plasma Assisted Reaction Chemical Ionization (PARCI) [138–

140]. In this strategy a plasma is utilized to breakdown compounds and to convert them to element- specific gas-phase neutrals which are then ionized using chemical ionization methods in the plasma afterglow. The approach provides fundamental improvements offered by tunability of chemical ionization and opens a path for resolving isobaric interferences by facile implementation of the ionization approach on molecular mass spectrometers with advanced ion manipulation capabilities.

17

Recent investigations have demonstrated the effectiveness of the general PARCI approach for detection of halogens in gaseous samples. In particular, the Jorabchi group has developed a

GC-PARCI-MS instrument utilizing a low-pressure helium microwave induced plasma (MIP) for breakdown of compounds prior to chemical ionization in negative mode [138–140]. Notably, GC-

MIP-PARCI-MS offers significantly improved sensitivity and limits of detection for chlorine and fluorine compared to GC-ICP-MS methods for fluorine and chlorine. There are however, two major drawbacks to GC-MIP-PARCI-MS, both stemming from the use of the low-pressure MIP.

The first is a practical problem: operation at low-pressure makes it difficult to couple to mass spectrometers as the MS has to be vented and the low-pressure MIP-PARCI interface installed in order to attain the needed pressures. The second is a fundamental problem: due to the relatively low power of the helium MIP, this plasma is not able to maintain breakdown efficiency in the presence of solvent, compromising its operation when coupled to liquid chromatography.

Accordingly, this dissertation focuses on developing a PARCI-based method for analysis of liquid samples and interfacing with liquid chromatography. This is accomplished by using an inductively coupled plasma for breakdown of compounds separated by liquid chromatography with a new atmospheric-pressure reaction interface that allows for efficient chemical ionization in the ICP afterglow. Novel ionization chemistries are explored to enhance the sensitivity of F and Cl elemental analysis.

1.6. Dissertation Outline

Chapter II has been reprinted with permission (See Appendix) from Lesniewski, J.E.,

McMahon, W.P., Zheng, K., Wang, H., Badiei, H., and Jorabchi, K.: Atmospheric pressure plasma assisted reaction chemical ionization for analysis of chlorinated compounds separated by liquid

18 chromatography. J. Anal. At. Spectrom. 32, 1757-1765 (2017). Copyright 2017 Royal Society of

Chemistry. This work provides a framework for the use of an inductively coupled plasma with a new interface to carry out post-plasma chemical reactions for chlorine quantitation.

Chapter III has been modified, integrated with supplementary information, and reprinted with permission (see Appendix) from Lesniewski, J.E., McMahon, W.P., and Jorabchi, K.:

Mechanistic insights into chloride ion detection from the atmospheric-pressure afterglow of an argon inductively coupled plasma. J. Anal. At. Spectrom. 33, 1981-1992 (2018). Copyright 2018

Royal Society of Chemistry. This work expands on the approach described in chapter II by developing methodologies for elucidating mechanisms of ionization and measuring physical transport parameters of the atmospheric pressure reaction interface.

Chapter IV has been modified, integrated with supplementary information, and reprinted with permission (see Appendix) from Lesniewski, J.E., Zheng, K., Lecchi, P., Dain, D., and

Jorabchi, K.: High-Sensitivity Elemental Mass Spectrometry of Fluorine by Ionization in Plasma

Afterglow. Anal. Chem. 91, 3773-3777 (2019). Copyright 2019 American Society of Chemistry.

This work further extends the concepts developed in chapters II and III by introducing new ionization reactions allowing for positive mode ionization of nonmetals. These reactions are applied to fluorine ionization, and a real-world sample analysis of fluoride in infant formula.

Chapter V provides a unified summary of preceding chapters and explores preliminary findings in extension of the methods to other nonmetal elements: P and S. Data presented are investigated with the aim of eventual development of simultaneous and sensitive compound- independent quantitation of F, Cl, P, and S using atmospheric pressure post-plasma chemical ionization.

19

1.7. References

1. Carter, S., Fisher, A., Garcia, R., Gibson, B., Lancaster, S., Marshall, J., Whiteside, I.:

Atomic spectrometry update. Review of advances in the analysis of metals, chemicals and functional materials. J. Anal. At. Spectrom. 30, 2249–2294 (2015). https://doi.org/10.1039/C5JA90045J

2. Taylor, A., Barlow, N., Day, M.P., Hill, S., Martin, N., Patriarca, M.: Atomic Spectrometry

Update: review of advances in the analysis of clinical and biological materials, foods and beverages. J. Anal. At. Spectrom. 33, 338–382 (2018). https://doi.org/10.1039/C8JA90005A

3. Bacon, J.R., Butler, O.T., Cairns, W.R.L., Cook, J.M., Mertz-Kraus, R., Tyson, J.F.:

Atomic Spectrometry Update – a review of advances in environmental analysis. J. Anal. At.

Spectrom. 34, 9–58 (2019). https://doi.org/10.1039/C8JA90044B

4. Feldmann, J., Raab, A., Krupp, E.M.: Importance of ICPMS for speciation analysis is changing: future trends for targeted and non-targeted element speciation analysis. Anal. Bioanal.

Chem. 1–7 (2017). https://doi.org/10.1007/s00216-017-0502-8

5. Flores, E.M.M., Mello, P.A., Krzyzaniak, S.R., Cauduro, V.H., Picoloto, R.S.: Challenges and trends for halogen determination by inductively coupled plasma mass spectrometry: A review.

Rapid Commun. Mass Spectrom. 34, e8727 (2020). https://doi.org/10.1002/rcm.8727

6. Gubal, A., Chuchina, V., Sorokina, A., Solovyev, N., Ganeev, A.: Mass Spectrometry-

Based Techniques for Direct Quantification of High Ionization Energy Elements in Solid

Materials—Challenges and Perspectives. Mass Spectrom. Rev. n/a, (2020). https://doi.org/10.1002/mas.21643

20

7. Niu, H., Houk, R.S.: Fundamental aspects of ion extraction in inductively coupled plasma mass spectrometry. Spectrochim. Acta Part B At. Spectrosc. 51, 779–815 (1996). https://doi.org/10.1016/0584-8547(96)01506-6

8. Gallidabino, M.D., Irlam, R.C., Salt, M.C., O’Donnell, M., Beardah, M.S., Barron, L.P.:

Targeted and non-targeted forensic profiling of black powder substitutes and gunshot residue using gradient ion chromatography – high resolution mass spectrometry (IC-HRMS). Anal. Chim. Acta.

1072, 1–14 (2019). https://doi.org/10.1016/j.aca.2019.04.048

9. Oberacher, H., Arnhard, K.: Current status of non-targeted liquid chromatography-tandem mass spectrometry in forensic toxicology. TrAC Trends Anal. Chem. 84, 94–105 (2016). https://doi.org/10.1016/j.trac.2015.12.019

10. Fu, Y., Zhao, C., Lu, X., Xu, G.: Nontargeted screening of chemical contaminants and illegal additives in food based on liquid chromatography–high resolution mass spectrometry.

TrAC Trends Anal. Chem. 96, 89–98 (2017). https://doi.org/10.1016/j.trac.2017.07.014

11. Knolhoff, A.M., Croley, T.R.: Non-targeted screening approaches for contaminants and adulterants in food using liquid chromatography hyphenated to high resolution mass spectrometry.

J. Chromatogr. A. 1428, 86–96 (2016). https://doi.org/10.1016/j.chroma.2015.08.059

12. Lachenmeier, D.W., Humpfer, E., Fang, F., Schütz, B., Dvortsak, P., Sproll, C., Spraul,

M.: NMR-Spectroscopy for Nontargeted Screening and Simultaneous Quantification of Health-

Relevant Compounds in Foods: The Example of Melamine. J. Agric. Food Chem. 57, 7194–7199

(2009). https://doi.org/10.1021/jf902038j

21

13. Lorenc, W., Hanć, A., Sajnóg, A., Barałkiewicz, D.: Lc/Icp-Ms and Complementary

Techniques in Bespoke and Nontargeted Speciation Analysis of Elements in Food Samples. Mass

Spectrom. Rev. 00, 1–19 (2020). https://doi.org/10.1002/mas.21662

14. Mollerup, C.B., Dalsgaard, P.W., Mardal, M., Linnet, K.: Targeted and non-targeted drug screening in whole blood by UHPLC-TOF-MS with data-independent acquisition. Drug Test.

Anal. 9, 1052–1061 (2017). https://doi.org/10.1002/dta.2120

15. Scheidweiler, K.B., Jarvis, M.J.Y., Huestis, M.A.: Nontargeted SWATH acquisition for identifying 47 synthetic cannabinoid metabolites in human urine by liquid chromatography-high- resolution tandem mass spectrometry. Anal. Bioanal. Chem. 407, 883–897 (2015). https://doi.org/10.1007/s00216-014-8118-8

16. Hilton, D.C., Jones, R.S., Sjödin, A.: A method for rapid, non-targeted screening for environmental contaminants in household dust. J. Chromatogr. A. 1217, 6851–6856 (2010). https://doi.org/10.1016/j.chroma.2010.08.039

17. Hollender, J., Schymanski, E.L., Singer, H.P., Ferguson, P.L.: Nontarget Screening with

High Resolution Mass Spectrometry in the Environment: Ready to Go? Environ. Sci. Technol. 51,

11505–11512 (2017). https://doi.org/10.1021/acs.est.7b02184

18. Hug, C., Ulrich, N., Schulze, T., Brack, W., Krauss, M.: Identification of novel micropollutants in wastewater by a combination of suspect and nontarget screening. Environ.

Pollut. 184, 25–32 (2014). https://doi.org/10.1016/j.envpol.2013.07.048

19. Cho, W.-S., Han, B.S., Nam, K.T., Park, K., Choi, M., Kim, S.H., Jeong, J., Jang, D.D.:

Carcinogenicity study of 3-monochloropropane-1,2-diol in Sprague–Dawley rats. Food Chem.

Toxicol. 46, 3172–3177 (2008). https://doi.org/10.1016/j.fct.2008.07.003

22

20. Velíšek, J., Davidek, J., Hajšlová, J., Kubelka, V., Janíček, G., Mánková, B.: Chlorohydrins in protein hydrolysates. Z. Für Lebensm.-Unters. Forsch. 167, 241–244 (1978). https://doi.org/10.1007/BF01135595

21. Wenzl, T., Lachenmeier, D.W., Gökmen, V.: Analysis of heat-induced contaminants

(acrylamide, chloropropanols and furan) in carbohydrate-rich food. Anal. Bioanal. Chem. 389, 119

(2007). https://doi.org/10.1007/s00216-007-1459-9

22. Timmerman, P., Blech, S., White, S., Green, M., Delatour, C., McDougall, S., Mannens,

G., Smeraglia, J., Williams, S., Young, G.: Best practices for metabolite quantification in drug development: updated recommendation from the European Bioanalysis Forum. Bioanalysis. 8,

1297–1305 (2016). https://doi.org/10.4155/bio-2016-0103

23. Chen, Y., Monshouwer, M., Fitch, W.L.: Analytical Tools and Approaches for Metabolite

Identification in Early Drug Discovery. Pharm. Res. 24, 248–257 (2007). https://doi.org/10.1007/s11095-006-9162-7

24. Gammelgaard, B., Hansen, H.R., Stürup, S., Møller, C.: The use of inductively coupled plasma mass spectrometry as a detector in drug metabolism studies. Expert Opin. Drug Metab.

Toxicol. 4, 1187–1207 (2008). https://doi.org/10.1517/17425255.4.9.1187

25. Luffer-Atlas, D.: The early estimation of circulating drug metabolites in humans. Expert

Opin. Drug Metab. Toxicol. 8, 985–997 (2012). https://doi.org/10.1517/17425255.2012.693159

26. Makarov, A., Scigelova, M.: Coupling liquid chromatography to Orbitrap mass spectrometry. J. Chromatogr. A. 1217, 3938–3945 (2010). https://doi.org/10.1016/j.chroma.2010.02.022

23

27. Makarov, A., Denisov, E., Kholomeev, A., Balschun, W., Lange, O., Strupat, K., Horning,

S.: Performance Evaluation of a Hybrid Linear Ion Trap/Orbitrap Mass Spectrometer. Anal. Chem.

78, 2113–2120 (2006). https://doi.org/10.1021/ac0518811

28. Makarov, A., Denisov, E., Kholomeev, A., Balschun, W., Lange, O., Strupat, K., Horning,

S.: Performance Evaluation of a Hybrid Linear Ion Trap/Orbitrap Mass Spectrometer. Anal. Chem.

78, 2113–2120 (2006). https://doi.org/10.1021/ac0518811

29. Zubarev, R.A., Makarov, A.: Orbitrap Mass Spectrometry. Anal. Chem. 85, 5288–5296

(2013). https://doi.org/10.1021/ac4001223

30. Scigelova, M., Makarov, A.: Orbitrap Mass Analyzer – Overview and Applications in

Proteomics. PROTEOMICS. 6, 16–21 (2006). https://doi.org/10.1002/pmic.200600528

31. Hu, Q., Noll, R.J., Li, H., Makarov, A., Hardman, M., Cooks, R.G.: The Orbitrap: a new mass spectrometer. J. Mass Spectrom. 40, 430–443 (2005). https://doi.org/10.1002/jms.856

32. Michalski, A., Damoc, E., Hauschild, J.-P., Lange, O., Wieghaus, A., Makarov, A.,

Nagaraj, N., Cox, J., Mann, M., Horning, S.: Mass Spectrometry-based Proteomics Using Q

Exactive, a High-performance Benchtop Quadrupole Orbitrap Mass Spectrometer. Mol. Cell.

Proteomics. 10, (2011). https://doi.org/10.1074/mcp.M111.011015

33. Covey, T.R., Thomson, B.A., Schneider, B.B.: Atmospheric pressure ion sources. Mass

Spectrom. Rev. 28, 870–897 (2009). https://doi.org/10.1002/mas.20246

34. Gong, Y., Chen, J., Shi, Y., Lim, H.-K., Weng, N., Salter, R.: Standard-Free Bioanalytical

Approach for Absolute Quantitation of Drug Metabolites Utilizing Biosynthesis of Reciprocal

Radio and Stable Isotopologues and Its Application. Anal. Chem. 89, 8399–8404 (2017). https://doi.org/10.1021/acs.analchem.7b01830

24

35. Schymanski, E.L., Singer, H.P., Slobodnik, J., Ipolyi, I.M., Oswald, P., Krauss, M.,

Schulze, T., Haglund, P., Letzel, T., Grosse, S., Thomaidis, N.S., Bletsou, A., Zwiener, C., Ibáñez,

M., Portolés, T., de Boer, R., Reid, M.J., Onghena, M., Kunkel, U., Schulz, W., Guillon, A.,

Noyon, N., Leroy, G., Bados, P., Bogialli, S., Stipaničev, D., Rostkowski, P., Hollender, J.: Non- target screening with high-resolution mass spectrometry: critical review using a collaborative trial on water analysis. Anal. Bioanal. Chem. 407, 6237–6255 (2015). https://doi.org/10.1007/s00216-

015-8681-7

36. Ibáñez, M., Sancho, J.V., Hernández, F., McMillan, D., Rao, R.: Rapid non-target screening of organic pollutants in water by ultraperformance liquid chromatography coupled to time-of-light mass spectrometry. TrAC Trends Anal. Chem. 27, 481–489 (2008). https://doi.org/10.1016/j.trac.2008.03.007

37. Meng, C.-K., Zweigenbaum, J., Fürst, P., Blanke, E.: Finding and Confirming Nontargeted

Pesticides Using GC/MS, LC/Quadrupole-Time-of-Flight MS, and Databases. J. AOAC Int. 93,

703–711 (2010). https://doi.org/10.1093/jaoac/93.2.703

38. Fernández, S.D., Sugishama, N., Encinar, J.R., Sanz-Medel, A.: Triple Quad ICPMS

(ICPQQQ) as a New Tool for Absolute Quantitative Proteomics and Phosphoproteomics. Anal.

Chem. 84, 5851–5857 (2012). https://doi.org/10.1021/ac3009516

39. Calderón-Celis, F., Sanz-Medel, A., Ruiz Encinar, J.: Universal absolute quantification of biomolecules using element mass spectrometry and generic standards. Chem. Commun. 54, 904–

907 (2018). https://doi.org/10.1039/C7CC09059E

40. O’Hagan, D.: Fluorine in health care: Organofluorine containing blockbuster drugs. J.

Fluor. Chem. 131, 1071–1081 (2010). https://doi.org/10.1016/j.jfluchem.2010.03.003

25

41. Secretariat of the Stockholm Convention: The Stockholm Convention on Persistent

Organic Pollutants, http://chm.pops.int/

42. US Food & Drug Administration: Pesticide Residue Monitoring Program, https://www.fda.gov/Food/FoodborneIllnessContaminants/Pesticides/ucm582707.htm

43. Yu, J., Nickerson, A., Li, Y., Fang, Y., Strathmann, T.J.: Fate of per- and polyfluoroalkyl substances (PFAS) during hydrothermal liquefaction of municipal wastewater treatment sludge.

Environ. Sci. Water Res. Technol. 6, 1388–1399 (2020). https://doi.org/10.1039/C9EW01139K

44. Stoiber, T., Evans, S., Naidenko, O.V.: Disposal of products and materials containing per- and polyfluoroalkyl substances (PFAS): A cyclical problem. Chemosphere. 260, 127659 (2020). https://doi.org/10.1016/j.chemosphere.2020.127659

45. Singh, R.K., Fernando, S., Baygi, S.F., Multari, N., Thagard, S.M., Holsen, T.M.:

Breakdown Products from Perfluorinated Alkyl Substances (PFAS) Degradation in a Plasma-

Based Water Treatment Process. Environ. Sci. Technol. 53, 2731–2738 (2019). https://doi.org/10.1021/acs.est.8b07031

46. Lavonen, E.E., Gonsior, M., Tranvik, L.J., Schmitt-Kopplin, P., Köhler, S.J.: Selective

Chlorination of Natural Organic Matter: Identification of Previously Unknown Disinfection

Byproducts. Environ. Sci. Technol. 47, 2264–2271 (2013). https://doi.org/10.1021/es304669p

47. Szczuka, A., Parker, K.M., Harvey, C., Hayes, E., Vengosh, A., Mitch, W.A.: Regulated and unregulated halogenated disinfection byproduct formation from chlorination of saline groundwater. Water Res. 122, 633–644 (2017). https://doi.org/10.1016/j.watres.2017.06.028

48. Daiber, E.J., DeMarini, D.M., Ravuri, S.A., Liberatore, H.K., Cuthbertson, A.A.,

Thompson-Klemish, A., Byer, J.D., Schmid, J.E., Afifi, M.Z., Blatchley, E.R., Richardson, S.D.:

26

Progressive Increase in Disinfection Byproducts and Mutagenicity from Source to Tap to

Swimming Pool and Spa Water: Impact of Human Inputs. Environ. Sci. Technol. 50, 6652–6662

(2016). https://doi.org/10.1021/acs.est.6b00808

49. Schaider, L.A., Balan, S.A., Blum, A., Andrews, D.Q., Strynar, M.J., Dickinson, M.E.,

Lunderberg, D.M., Lang, J.R., Peaslee, G.F.: Fluorinated Compounds in U.S. Fast Food

Packaging. Environ. Sci. Technol. Lett. 4, 105–111 (2017). https://doi.org/10.1021/acs.estlett.6b00435

50. Li, S., Huang, J., Chen, Z., Chen, G., Lai, Y.: A review on special wettability textiles: theoretical models, fabrication technologies and multifunctional applications. J. Mater. Chem. A.

5, 31–55 (2017). https://doi.org/10.1039/C6TA07984A

51. Schultes, L., Peaslee, G.F., Brockman, J.D., Majumdar, A., McGuinness, S.R., Wilkinson,

J.T., Sandblom, O., Ngwenyama, R.A., Benskin, J.P.: Total Fluorine Measurements in Food

Packaging: How Do Current Methods Perform? Environ. Sci. Technol. Lett. 6, 73–78 (2019). https://doi.org/10.1021/acs.estlett.8b00700

52. Barzen-Hanson, K.A., Roberts, S.C., Choyke, S., Oetjen, K., McAlees, A., Riddell, N.,

McCrindle, R., Ferguson, P.L., Higgins, C.P., Field, J.A.: Discovery of 40 Classes of Per- and

Polyfluoroalkyl Substances in Historical Aqueous Film-Forming Foams (AFFFs) and AFFF-

Impacted Groundwater. Environ. Sci. Technol. 51, 2047–2057 (2017). https://doi.org/10.1021/acs.est.6b05843

53. Houtz, E.F., Higgins, C.P., Field, J.A., Sedlak, D.L.: Persistence of Perfluoroalkyl Acid

Precursors in AFFF-Impacted Groundwater and Soil. Environ. Sci. Technol. 47, 8187–8195

(2013). https://doi.org/10.1021/es4018877

27

54. Poothong, S., Thomsen, C., Padilla-Sanchez, J.A., Papadopoulou, E., Haug, L.S.:

Distribution of Novel and Well-Known Poly- and Perfluoroalkyl Substances (PFASs) in Human

Serum, Plasma, and Whole Blood. Env. Sci Technol. 51, 13388–13396 (2017). https://doi.org/10.1021/acs.est.7b03299

55. Negreira, N., Regueiro, J., López de Alda, M., Barceló, D.: Degradation of the anticancer drug erlotinib during water chlorination: Non-targeted approach for the identification of transformation products. Water Res. 85, 103–113 (2015). https://doi.org/10.1016/j.watres.2015.08.005

56. Murphy, C.D.: Microbial degradation of fluorinated drugs: biochemical pathways, impacts on the environment and potential applications. Appl. Microbiol. Biotechnol. 100, 2617–2627

(2016). https://doi.org/10.1007/s00253-016-7304-3

57. Liu, J., Mejia Avendaño, S.: Microbial degradation of polyfluoroalkyl chemicals in the environment: A review. Environ. Int. 61, 98–114 (2013). https://doi.org/10.1016/j.envint.2013.08.022

58. Brzezicki, J.M., Andersen, M.E., Cranmer, B.K., Tessari, J.D.: Quantitative Identification of Atrazine and its Chlorinated Metabolites in Plasma. J. Anal. Toxicol. 27, 569–573 (2003). https://doi.org/10.1093/jat/27.8.569

59. Krol, G.J., Noe, A.J., Beermann, D.: Liquid Chromatographic Analysis of Ciprofloxacin and Ciprofloxacin Metabolites in Body Fluids. J. Liq. Chromatogr. 9, 2897–2919 (1986). https://doi.org/10.1080/01483918608074158

60. Müller, K., Faeh, C., Diederich, F.: Fluorine in Pharmaceuticals: Looking Beyond

Intuition. Science. 317, 1881–1886 (2007). https://doi.org/10.1126/science.1131943

28

61. Resano, M., Flórez, M.R., García-Ruiz, E.: Progress in the determination of metalloids and non-metals by means of high-resolution continuum source atomic or molecular absorption spectrometry. A critical review. Anal. Bioanal. Chem. 406, 2239–2259 (2014). https://doi.org/10.1007/s00216-013-7522-9

62. Houseaux, J., Mermet, J.-M.: Use of a charge-coupled device detector in the 120–190 nm range in axially-viewed inductively coupled plasma atomic emission spectrometry. J. Anal. At.

Spectrom. 15, 979–982 (2000). https://doi.org/10.1039/B003626I

63. LaFreniere, B.R., Houk, R.S., Fassel, V.A.: Direct detection of vacuum ultraviolet radiation through an optical sampling orifice: analytical figures of merit for the nonmetals, metalloids, and selected metals by inductively coupled plasma atomic emission spectrometry.

Anal. Chem. 59, 2276–2282 (1987). https://doi.org/10.1021/ac00145a015

64. Wheal, M.S., Palmer, L.T.: Chloride analysis of botanical samples by ICP-OES. J. Anal.

At. Spectrom. 25, 1946–1952 (2010). https://doi.org/10.1039/C0JA00059K

65. Flores, É.M.M., Mesko, M.F., Moraes, D.P., Pereira, J.S.F., Mello, P.A., Barin, J.S.,

Knapp, G.: Determination of Halogens in Coal after Digestion Using the Microwave-Induced

Combustion Technique. Anal. Chem. 80, 1865–1870 (2008). https://doi.org/10.1021/ac8000836

66. Fry, R.C., Northway, S.J., Brown, R.M., Hughes, S.K.: Atomic fluorine spectra in the argon inductively coupled plasma. Anal. Chem. 52, 1716–1722 (1980). https://doi.org/10.1021/ac50061a042

67. Kovács, M., Magarini, R., Halmos, P.: Automated indirect fluoride determination by ICP-

OES using discontinuous-flow analysis. Toxicol. Environ. Chem. 91, 1217–1227 (2009). https://doi.org/10.1080/02772240802631907

29

68. Morés, S., Monteiro, G.C., Santos, F. da S., Carasek, E., Welz, B.: Determination of fluorine in tea using high-resolution molecular absorption spectrometry with electrothermal vaporization of the calcium mono-fluoride CaF. Talanta. 85, 2681–2685 (2011). https://doi.org/10.1016/j.talanta.2011.08.044

69. Welz, B., Lepri, F.G., Araujo, R.G.O., Ferreira, S.L.C., Huang, M.-D., Okruss, M., Becker-

Ross, H.: Determination of phosphorus, sulfur and the halogens using high-temperature molecular absorption spectrometry in flames and furnaces—A review. Anal. Chim. Acta. 647, 137–148

(2009). https://doi.org/10.1016/j.aca.2009.06.029

70. Machyňák, Ľ., Čacho, F., Němeček, M., Beinrohr, E.: Determination of trace concentrations of chlorine in aqueous solutions by high-resolution continuum source graphite furnace molecular absorption spectrometry. Spectrochim. Acta Part B At. Spectrosc. 125, 140–

145 (2016). https://doi.org/10.1016/j.sab.2016.10.006

71. Borges, A.R., Duarte, Á.T., Potes, M. da L., Silva, M.M., Vale, M.G.R., Welz, B.: Fluorine in eye shadow: Development of method using high-resolution continuum source graphite furnace molecular absorption spectrometry via calcium mono-fluoride with direct solid sample introduction. Microchem. J. 124, 410–415 (2016). https://doi.org/10.1016/j.microc.2015.09.025

72. Qin, Z., McNee, D., Gleisner, H., Raab, A., Kyeremeh, K., Jaspars, M., Krupp, E., Deng,

H., Feldmann, J.: Fluorine Speciation Analysis Using Reverse Phase Liquid Chromatography

Coupled Off-Line to Continuum Source Molecular Absorption Spectrometry (CS-MAS):

Identification and Quantification of Novel Fluorinated Organic Compounds in Environmental and

Biological Samples. Anal. Chem. 84, 6213–6219 (2012). https://doi.org/10.1021/ac301201y

30

73. Gawor, A., Konopka, A., Ruszczyńska, A., Pączek, L., Gajewski, Z., Bulska, E.: Molecular absorption and mass spectrometry for complementary analytical study of fluorinated drugs in animal organisms. J. Anal. At. Spectrom. 35, 1840–1847 (2020). https://doi.org/10.1039/D0JA00126K

74. Krüger, M., Huang, M.-D., Becker-Roß, H., Florek, S., Ott, I., Gust, R.: Quantification of the fluorine containing drug 5-fluorouracil in cancer cells by GaF molecular absorption via high- resolution continuum source molecular absorption spectrometry. Spectrochim. Acta Part B At.

Spectrosc. 69, 50–55 (2012). https://doi.org/10.1016/j.sab.2012.02.004

75. Ozbek, N., Akman, S.: Solid sampling determination of total fluorine in baby food samples by high-resolution continuum source graphite furnace molecular absorption spectrometry. Food

Chem. 211, 180–184 (2016). https://doi.org/10.1016/j.foodchem.2016.05.044

76. Tang, Z., Zhou, R., Hao, Z., Zhang, W., Li, Q., Zeng, Q., Li, X., Zeng, X., Lu, Y.:

Determination of fluorine in copper ore using laser-induced breakdown spectroscopy assisted by the SrF molecular emission band. J. Anal. At. Spectrom. 35, 754–761 (2020). https://doi.org/10.1039/C9JA00407F

77. Zhang, W., Zhou, R., Yang, P., Liu, K., Yan, J., Gao, P., Tang, Z., Li, X., Lu, Y., Zeng,

X.: Determination of chlorine with radical emission using laser-induced breakdown spectroscopy coupled with partial least square regression. Talanta. 198, 93–96 (2019). https://doi.org/10.1016/j.talanta.2019.01.102

78. Tang, Z., Hao, Z., Zhou, R., Li, Q., Liu, K., Zhang, W., Yan, J., Wei, K., Li, X.: Sensitive analysis of fluorine and chlorine elements in water solution using laser-induced breakdown

31 spectroscopy assisted with molecular synthesis. Talanta. 121784 (2020). https://doi.org/10.1016/j.talanta.2020.121784

79. Martino, R., Gilard, V., Desmoulin, F., Malet-Martino, M.: Fluorine-19 or phosphorus-31

NMR spectroscopy: A suitable analytical technique for quantitative in vitro metabolic studies of fluorinated or phosphorylated drugs. J. Pharm. Biomed. Anal. 38, 871–891 (2005). https://doi.org/10.1016/j.jpba.2005.01.047

80. Gee, C.T., Arntson, K.E., Urick, A.K., Mishra, N.K., Hawk, L.M.L., Wisniewski, A.J.,

Pomerantz, W.C.K.: Protein-observed 19 F-NMR for fragment screening, affinity quantification and druggability assessment. Nat. Protoc. 11, 1414–1427 (2016). https://doi.org/10.1038/nprot.2016.079

81. Lim, H.-S., Lee, S.-G.: Quantitative Analysis of Chloride by Chlorine-35 NMR

Spectroscopy. Bull. Korean Chem. Soc. 27, 972–973 (2006). https://doi.org/10.5012/BKCS.2006.27.7.972

82. McDonough, C.A., Guelfo, J.L., Higgins, C.P.: Measuring total PFASs in water: The tradeoff between selectivity and inclusivity. Curr. Opin. Environ. Sci. Health. 7, 13–18 (2019). https://doi.org/10.1016/j.coesh.2018.08.005

83. Espina, R., Yu, L., Wang, J., Tong, Z., Vashishtha, S., Talaat, R., Scatina, J., Mutlib, A.:

Nuclear Magnetic Resonance Spectroscopy as a Quantitative Tool To Determine the

Concentrations of Biologically Produced Metabolites: Implications in Metabolites in Safety

Testing. Chem. Res. Toxicol. 22, 299–310 (2009). https://doi.org/10.1021/tx800251p

84. Holzgrabe, U.: Quantitative NMR spectroscopy in pharmaceutical applications. Prog.

Nucl. Magn. Reson. Spectrosc. 57, 229–240 (2010). https://doi.org/10.1016/j.pnmrs.2010.05.001

32

85. Howe, P.W.A.: Recent developments in the use of fluorine NMR in synthesis and characterisation. Prog. Nucl. Magn. Reson. Spectrosc. 118–119, 1–9 (2020). https://doi.org/10.1016/j.pnmrs.2020.02.002

86. Boeszoermenyi, A., Ogórek, B., Jain, A., Arthanari, H., Wagner, G.: The precious fluorine on the ring: fluorine NMR for biological systems. J. Biomol. NMR. 74, 365–379 (2020). https://doi.org/10.1007/s10858-020-00331-z

87. Orhan, H., Commandeur, J.N.M., Sahin, G., Aypar, U., Sahin, A., Vermeulen, N.P.E.: Use of 19F-nuclear magnetic resonance and -electron capture detection in the quantitative analysis of fluorine-containing metabolites in urine of sevoflurane-anaesthetized patients. Xenobiotica Fate Foreign Compd. Biol. Syst. 34, 301–316 (2004). https://doi.org/10.1080/716494157

88. Lambert, J.B., Schilif, Wojciech.: Chlorine NMR studies of ionized and associated silyl perchlorates. J. Am. Chem. Soc. 110, 6364–6367 (1988). https://doi.org/10.1021/ja00227a015

89. Hu, H., Katyayan, K.K., Czeskis, B.A., Perkins, E.J., Kulanthaivel, P.: Comparison between Radioanalysis and 19F Nuclear Magnetic Resonance Spectroscopy in the Determination of Mass Balance, Metabolism, and Distribution of Pefloxacin. Drug Metab. Dispos. 45, 399–408

(2017). https://doi.org/10.1124/dmd.116.073809

90. Montaser, A., Golightly, D.W.: Inductively Coupled Plasmas in Analytical Atomic

Spectrometry. VCH Publishers (1987)

91. Montaser, A.: Inductively Coupled Plasma Mass Spectrometry. John Wiley & Sons (1998)

33

92. Baglan, N., Cossonnet, C., Pitet, P., Cavadore, D., Exmelin, L., Berard, P.: On the Use of

ICP-MS for Measuring Plutonium in Urine. J. Radioanal. Nucl. Chem. 243, 397–401 (2004). https://doi.org/10.1023/a:1016013811974

93. Hu, Y., Vanhaecke, F., Moens, L., Dams, R.: Semi-quantitative panoramic analysis of industrial samples by inductively coupled plasma mass spectrometry. Anal. Chim. Acta. 355, 105–

111 (1997). https://doi.org/10.1016/S0003-2670(97)00496-0

94. Todolí, J.-L., Hernandis, V., Canals, A., Mermet, J.-M.: Comparison of characteristics and limits of detection of pneumatic micronebulizers and a conventional nebulizer operating at low uptake rates in ICP-AES. J. Anal. At. Spectrom. 14, 1289–1295 (1999). https://doi.org/10.1039/A900598F

95. Jorabchi, K., Kahen, K., Gray, C., Montaser, A.: In Situ and Characterization of Aerosol Droplets in an Inductively Coupled Plasma. Anal. Chem. 77, 1253–1260 (2005). https://doi.org/10.1021/ac048576k

96. Smith, D.D., Browner, R.F.: Measurement of aerosol transport efficiency in atomic spectrometry. Anal. Chem. 54, 533–537 (1982). https://doi.org/10.1021/ac00240a041

97. Browner, R.F., Canals, A., Hernandis, V.: Effect of analyte and solvent transport on signal intensities in inductively coupled plasma atomic emission spectrometry. Spectrochim. Acta Part B

At. Spectrosc. 47, 659–673 (1992). https://doi.org/10.1016/0584-8547(92)80060-T

98. Nam, S.-H., Lim, J.-S., Montaser, A.: High-efficiency nebulizer for argon inductively coupled plasma mass spectrometry. J. Anal. At. Spectrom. 9, 1357–1362 (1994). https://doi.org/10.1039/JA9940901357

34

99. Todolí, J.-L., Mermet, J.-M.: Evaluation of a direct injection high-efficiency nebulizer

(DIHEN) by comparison with a high-efficiency nebulizer (HEN) coupled to a cyclonic spray chamber as a liquid sample introduction system for ICP-AES. J. Anal. At. Spectrom. 16, 514–520

(2001). https://doi.org/10.1039/B009430G

100. Bings, N.H., Orlandini von Niessen, J.O., Schaper, J.N.: Liquid sample introduction in inductively coupled plasma atomic emission and mass spectrometry — Critical review.

Spectrochim. Acta Part B At. Spectrosc. 100, 14–37 (2014). https://doi.org/10.1016/j.sab.2014.08.011

101. Mermet, J.-M., Todolí, J.-L.: Towards total-consumption pneumatic liquid micro-sample- introduction systems in ICP spectrochemistry. Anal. Bioanal. Chem. 378, 57–59 (2004). https://doi.org/10.1007/s00216-003-2368-1

102. Todolí, J.-L., Mermet, J.-M.: Effect of the spray chamber design on steady and transient acid interferences in inductively coupled plasma atomic emission spectrometry Presented at the

1999 Federation of Analytical Chemistry and Spectroscopy Societies (FACSS), Vancouver,

Canada. J. Anal. At. Spectrom. 15, 863–867 (2000). https://doi.org/10.1039/B000589O

103. Savtchenko, S., Badiei, H.R.: Shock wave properties in the interface region of an inductively coupled plasma mass spectrometer (ICP-MS). Presented at the 22nd International

Symposium on Plasma Chemistry , Antwerp, Belgium July 5 (2015)

104. Rumble, J.R.: CRC handbook of chemistry and physics. (2017)

105. Okamoto, Y.: Determination of fluorine in aqueous samples by electrothermal vaporisation inductively coupled plasma mass spectrometry (ETV-ICP-MS). J. Anal. At. Spectrom. 16, 539–

541 (2001). https://doi.org/10.1039/B101969O

35

106. Klencsár, B., Balcaen, L., Cuyckens, F., Lynen, F., Vanhaecke, F.: Development and validation of a novel quantification approach for gradient elution reversed phase high-performance liquid chromatography coupled to tandem ICP-mass spectrometry (RP-HPLC-ICP-MS/MS) and its application to diclofenac and its related compounds. Anal. Chim. Acta. 974, 43–53 (2017). https://doi.org/10.1016/j.aca.2017.04.030

107. Klencsár, B., Bolea-Fernandez, E., Flórez, M.R., Balcaen, L., Cuyckens, F., Lynen, F.,

Vanhaecke, F.: Determination of the total drug-related chlorine and bromine contents in human blood plasma using high performance liquid chromatography–tandem ICP-mass spectrometry

(HPLC–ICP-MS/MS). J. Pharm. Biomed. Anal. 124, 112–119 (2016). https://doi.org/10.1016/j.jpba.2016.02.019

108. Klencsár, B., Li, S., Balcaen, L., Vanhaecke, F.: High-performance liquid chromatography coupled to inductively coupled plasma – Mass spectrometry (HPLC-ICP-MS) for quantitative metabolite profiling of non-metal drugs. TrAC Trends Anal. Chem. 104, 118–134 (2018). https://doi.org/10.1016/j.trac.2017.09.020

109. Klencsár, B., Bolea-Fernandez, E., Flórez, M.R., Balcaen, L., Cuyckens, F., Lynen, F.,

Vanhaecke, F.: Determination of the total drug-related chlorine and bromine contents in human blood plasma using high performance liquid chromatography–tandem ICP-mass spectrometry

(HPLC–ICP-MS/MS). J. Pharm. Biomed. Anal. 124, 112–119 (2016). https://doi.org/10.1016/j.jpba.2016.02.019

110. Bolea-Fernandez, E., Balcaen, L., Resano, M., Vanhaecke, F.: Overcoming spectral overlap via inductively coupled plasma-tandem mass spectrometry (ICP-MS/MS). A tutorial review. J. Anal. At. Spectrom. (2017). https://doi.org/10.1039/C7JA00010C

36

111. Böting, K., Treu, S., Leonhard, P., Heiß, C., Bings, N.H.: First experimental proof of asymmetric charge transfer in ICP-MS/MS (ICP-QQQ-MS) through isotopically enriched oxygen as cell gas. J. Anal. At. Spectrom. 29, 578–582 (2014). https://doi.org/10.1039/C3JA50234A

112. Pröfrock, D., Leonhard, P., Wilbur, S., Prange, A.: Sensitive, simultaneous determination of P, S, Cl, Br and I containing pesticides in environmental samples by GC hyphenated with collision-cell ICP-MS. J. Anal. At. Spectrom. 19, 623–631 (2004). https://doi.org/10.1039/B310530J

113. Pröfrock, D., Prange, A.: Inductively Coupled Plasma–Mass Spectrometry (ICP-MS) for

Quantitative Analysis in Environmental and Life Sciences: A Review of Challenges, Solutions, and Trends. Appl. Spectrosc. 66, 843–868 (2012). https://doi.org/10.1366/12-06681

114. Bu, X., Wang, T., Hall, G.: Determination of halogens in organic compounds by high resolution inductively coupled plasma mass spectrometry (HR-ICP-MS). J. Anal. At. Spectrom.

18, 1443–1451 (2003). https://doi.org/10.1039/B306570G

115. Blades, M.W., Caughlin, B.L.: Excitation temperature and electron density in the inductively coupled plasma—aqueous vs organic solvent introduction. Spectrochim. Acta Part B

At. Spectrosc. 40, 579–591 (1985). https://doi.org/10.1016/0584-8547(85)80107-5

116. Barrett, P., Pruszkowska, E.: Use of organic solvents for inductively coupled plasma analyses. Anal. Chem. 56, 1927–1930 (1984). https://doi.org/10.1021/ac00275a038

117. Todoli, J.L., Canals, A., Hernandis, V.: Behaviour of a single-bore high-pressure pneumatic nebulizer operating with alcohols in inductively coupled plasma atomic emission spectrometry. J. Anal. At. Spectrom. 11, 949–956 (1996). https://doi.org/10.1039/JA9961100949

37

118. Klencsár, B., Balcaen, L., Cuyckens, F., Lynen, F., Vanhaecke, F.: Development and validation of a novel quantification approach for gradient elution reversed phase high-performance liquid chromatography coupled to tandem ICP-mass spectrometry (RP-HPLC-ICP-MS/MS) and its application to diclofenac and its related compounds. Anal. Chim. Acta. 974, 43–53 (2017). https://doi.org/10.1016/j.aca.2017.04.030

119. Grindlay, G., Mora, J., de Loos-Vollebregt, M., Vanhaecke, F.: A systematic study on the influence of carbon on the behavior of hard-to-ionize elements in inductively coupled plasma– mass spectrometry. Spectrochim. Acta Part B At. Spectrosc. 86, 42–49 (2013). https://doi.org/10.1016/j.sab.2013.05.002

120. Allain, Pierre., Jaunault, Laurent., Mauras, Yves., Mermet, J.Michel., Delaporte, Thierry.:

Signal enhancement of elements due to the presence of carbon-containing compounds in inductively coupled plasma mass spectrometry. Anal. Chem. 63, 1497–1498 (1991). https://doi.org/10.1021/ac00014a028

121. Montaser, A., Hoven, R.L.V., Barnes, R.M.: Mixed-Gas, Molecular-Gas, and Helium

Inductively Coupled Plasmas for Analytical Atomic Spectrometry: A Critical Review. C R C Crit.

Rev. Anal. Chem. 18, 45–103 (1987). https://doi.org/10.1080/10408348708542801

122. Heitkemper, D., Creed, J., Caruso, J.A.: Helium Microwave-Induced Plasma Mass

Spectrometric Detection for Reversed-Phase High-Performance Liquid Chromatography. J.

Chromatogr. Sci. 28, 175–181 (1990). https://doi.org/10.1093/chromsci/28.4.175

123. Lindner, H., Bogaerts, A.: Multi-element model for the simulation of inductively coupled plasmas: Effects of helium addition to the central gas stream. Spectrochim. Acta Part B At.

Spectrosc. 66, 421–431 (2011). https://doi.org/10.1016/j.sab.2011.04.007

38

124. Jorabchi, K., Montaser, A.: New coil geometries for the study of secondary discharge in atmospheric-pressure helium inductively coupled plasma mass spectrometry. Spectrochim. Acta

Part B At. Spectrosc. 59, 1471–1479 (2004). https://doi.org/10.1016/j.sab.2004.07.009

125. Jorabchi, K., McCormick, R., Levine, J.A., Liu, H., Nam, S.-H., Montaser, A.: High efficiency nebulization for helium inductively coupled plasma mass spectrometry. Spectrochim.

Acta Part B At. Spectrosc. 61, 945–950 (2006). https://doi.org/10.1016/j.sab.2006.08.003

126. Fulford, J.E., Quan, E.S.K.: Negative Ions in Inductively Coupled Plasma Mass

Spectrometry. Appl. Spectrosc. 42, 425–428 (1988)

127. Vickers, G.H., Wilson, D.A., Hieftje, G.M.: Detection of negative ions by inductively coupled plasma mass spectrometry. Anal. Chem. 60, 1808–1812 (1988). https://doi.org/10.1021/ac00168a031

128. Chtaib, M., Schmit, J.-P.: Negative-ion mode in inductively coupled plasma mass spectrometry. J. Anal. At. Spectrom. 3, 315–318 (1988). https://doi.org/10.1039/JA9880300315

129. Pupyshev, A.A., Surikov, V.T.: Application of negative ions in inductively coupled plasma-mass spectrometry. Spectrochim. Acta Part B At. Spectrosc. 59, 1021–1031 (2004). https://doi.org/10.1016/j.sab.2004.05.011

130. Blondel, C., Cacciani, P., Delsart, C., Trainham, R.: High-resolution determination of the electron affinity of fluorine and bromine using crossed ion and laser beams. Phys. Rev. A. 40,

3698–3701 (1989). https://doi.org/10.1103/PhysRevA.40.3698

131. Berzinsh, U., Gustafsson, M., Hanstorp, D., Klinkmüller, A., Ljungblad, U., Mårtensson-

Pendrill, A.-M.: Isotope shift in the electron affinity of chlorine. Phys. Rev. A. 51, 231–238 (1995). https://doi.org/10.1103/PhysRevA.51.231

39

132. Zhu, Y., Nakano, K., Shikamori, Y.: Analysis of Fluorine in Drinking Water by ICP-

QMS/QMS with an Octupole Reaction Cell. Anal. Sci. 33, 1279–1284 (2017). https://doi.org/10.2116/analsci.33.1279

133. Guo, W., Jin, L., Hu, S., Guo, Q.: Method Development for the Determination of Total

Fluorine in Foods by Tandem Inductively Coupled Plasma Mass Spectrometry with a Mass-Shift

Strategy. J. Agric. Food Chem. 65, 3406–3412 (2017). https://doi.org/10.1021/acs.jafc.7b00535

134. Jamari, N.L.A., Dohmann, J.F., Raab, A., Krupp, E.M., Feldmann, J.: Novel non-target analysis of fluorine compounds using ICPMS/MS and HPLC-ICPMS/MS. J. Anal. At. Spectrom.

32, 942–950 (2017). https://doi.org/10.1039/C7JA00051K

135. Jamari, N.L.A., Behrens, A., Raab, A., Krupp, E.M., Feldmann, J.: Plasma processes to detect fluorine with ICPMS/MS as [M–F]+: an argument for building a negative mode

ICPMS/MS. J. Anal. At. Spectrom. 33, 1304–1309 (2018). https://doi.org/10.1039/C8JA00050F

136. Jamari, N.L.A., Dohmann, J.F., Raab, A., Krupp, E.M., Feldmann, J.: Novel non-targeted analysis of perfluorinated compounds using fluorine-specific detection regardless of their ionisability (HPLC-ICPMS/MS-ESI-MS). Anal. Chim. Acta. 1053, 22–31 (2019). https://doi.org/10.1016/j.aca.2018.11.037

137. Guo, W., Lin, X., Jin, L., Hu, S.: Single quadrupole inductively coupled plasma-mass spectrometry for the measurement of fluorine in tea infusions and its health risk assessment. J.

Food Compos. Anal. 86, 103378 (2020). https://doi.org/10.1016/j.jfca.2019.103378

138. Wang, H., Lin, N., Kahen, K., Badiei, H., Jorabchi, K.: Plasma-Assisted Reaction

Chemical Ionization for Elemental Mass Spectrometry of Organohalogens. J. Am. Soc. Mass

Spectrom. 25, 692–695 (2014). https://doi.org/10.1007/s13361-013-0816-5

40

139. Lin, N., Wang, H., Kahen, K., Badiei, H., Jorabchi, K.: Gas Chromatography Plasma-

Assisted Reaction Chemical Ionization Mass Spectrometry for Quantitative Detection of Bromine in Organic Compounds. Anal. Chem. 86, 7954–7961 (2014). https://doi.org/10.1021/ac501964u

140. Wang, H., Minardi, C.S., Badiei, H., Kahen, K., Jorabchi, K.: High-sensitivity elemental ionization for quantitative detection of halogenated compounds. Analyst. 140, 8177–8185 (2015). https://doi.org/10.1039/C5AN01958C

41

Chapter II

Atmospheric Pressure Plasma Assisted Reaction Chemical Ionization for Analysis of

Chlorinated Compounds Separated by Liquid Chromatography*

2.1. Introduction

Inductively coupled plasma mass spectrometry (ICP-MS) has grown to become a robust and sensitive analytical method for elemental and isotopic analyses. Much of the success of Ar

ICP-MS is related to high ionization efficiencies (> 90%) for most elements within the 6000-8000

K plasma and positive ion extraction into the MS with minimal ion-molecule reactions in the sampling process [1]. This ionization and ion sampling approach, however, has encountered challenges in detection of non-metals, especially for low mass halogens. High ionization potentials of halogens lead to low ionization efficiencies in positive mode (e.g. 0.9% for Cl and 9x10-4% for

F) [1]. Space charge effects from Ar+ ions also reduce the transmission of low mass halogen ions into the MS. Moreover, isobaric interferences abundant at low mass (e.g. 34SH+ and 16O18OH+ for

35Cl+, and 36ArH+ and 36SH+ for 37Cl+) diminish the ability to detect low concentrations of these elements [2].

Recent technological advances have addressed some of the challenges outlined above. For example, reaction cells and triple quadrupole instruments have been utilized to react Cl+ ions with hydrogen and oxygen at low pressure, resulting in mass shift to m/z values free from isobaric

*This chapter has been reproduced from Lesniewski, J.E., McMahon, W.P., Zheng, K., Wang, H., Badiei, H., and Jorabchi, K.: Atmospheric pressure plasma assisted reaction chemical ionization for analysis of chlorinated compounds separated by liquid chromatography. J. Anal. At. Spectrom. 32, 1757-1765 (2017). with permission (see Appendix) from The Royal Society of Chemistry. The authors would like to acknowledge that this material is based upon work supported by the National Science Foundation (NSF) under CHE-1507304. The authors thank PerkinElmer Inc. for the loan of the mass spectrometer and the ICP generator used in these studies. The authors are grateful to Dr. Dickson Cheung and Dr. William Fisher of PerkinElmer for their assistance in setting up and maintaining the ICP generator. 42 interferences [3–6]. High resolution instruments have also been investigated to resolve the interfering ions for Cl isotopic measurements [7]. All of these approaches, however, improve signal-to-noise at the expense of absolute ion current and lead to more complex and costly instruments. Further, mass shift strategies limit multi-elemental capabilities within a single method. Space charge effects have been even more difficult to alleviate. For example, a recent implementation of an ion funnel for ICP-MS led to reduced analytical performance [8].

Fundamental improvement of the ion formation process holds promise to enhance sensitivity for low-mass high-ionization-potential elements. One approach has been development of helium plasmas where higher ionization temperatures yield enhanced positive ion formation while the low mass of helium reduces space charge effects [9]. However, limited success has been obtained in such approaches partly owing to the low gas temperatures of helium plasmas, leading to enhanced polyatomic interferences, increased matrix effects, and difficulties in liquid sample introduction [10–12].

Negative mode ICP-MS has been explored as a potential solution to the limitations of positive mode ionization [13–15]. Experiments and theoretical calculations have established that the negative ions observed using the conventional two-aperture ICP-MS ion sampling interface originate from ionization processes downstream of the supersonic gas expansion into vacuum [15,

16]. In other words, cooling of the gas is required for the negative ions to form in appreciable concentrations. Despite its promise, negative mode ICP-MS has not received much attention, perhaps because of meager improvements observed in early experiments relative to positive mode

ICP-MS and a lack of commercial instruments with electronics for negative mode ion detection.

43

To fully realize the potential of elemental analysis using negative ions we have recently developed an ionization method, termed plasma assisted reaction chemical ionization (PARCI). In

PARCI, analytes are subjected to plasma reactions, converting them to predictable species such as atoms and diatomic molecules [17–20]. The neutrals are then guided into an afterglow reaction area where lower temperatures enable formation of negative ions. This approach is in contrast with conventional elemental MS methods in which ions are typically sampled directly from the plasma and ion-neutral interactions are minimized. We have demonstrated the analytical utility of GC-

PARCI-MS for detection of Br-, Cl-, and F- with detection limits of 10 fg - 1 pg halogen on column using a low-pressure helium microwave induced plasma [18, 19]. Further, quantification of carbon via the production of CN- as reporter ions has proven to be successful with GC-PARCI-MS [20].

The low-pressure helium plasma for PARCI, however, has a number of shortcomings when coupled with liquid chromatography. Key among them, is the poor desolvation of liquid droplets during their residence time in the low-pressure helium plasma. In our earlier investigation, we utilized extensive desolvation along with a momentum separator to introduce liquid samples [17].

Major differences between response factors for organohalogens were observed because of sample introduction biases. Here, we report development of an ICP-based atmospheric pressure PARCI as an elemental detector of liquid chromatography. An atmospheric pressure interface is designed to couple PARCI with an atmospheric sampling mass spectrometer. We investigate the ion formation processes and demonstrate the high sensitivity of LC-PARCI and compound- independent response factors for Cl detection. We have selected chlorine as our test element because of growing interest in elemental chlorine analysis, owing to its presence in an increasing number of drugs, pesticides, and environmental contaminants [5, 21, 22].

44

2.2. Experimental Details

A schematic diagram of LC-PARCI-MS is shown in Figure 2.1. Liquid samples were introduced using a high efficiency nebulizer (HEN-170, Meinhard, Golden, CO) and a cyclonic spray chamber (Twister Cyclonic P/N 20-809-0365, Glass Expansion, Pocasset, MA) at a solvent flow rate of 50 µL/min supplied by a binary HPLC pump (SCL10AVP controller and dual

LC10AD pumps, Shimadzu, Colombia, MD). A second nebulizer-cyclonic spray chamber setup

(AR-30-1-FM01E nebulizer, Twister cyclonic P/N 20-809-0222 spray chamber, Glass Expansion,

Pocasset, MA) was used for independent introduction of makeup gas and reagents supplied by a syringe pump (Model 100, KD Scientific, Holliston, MA). Aerosol from both spray chambers was mixed using a tee and directed into the 1.5 mm injector of the ICP torch. A standalone ICP generator (NexION 2000, PerkinElmer Inc., Waltham, MA) with a built-in flow control module was used for plasma generation. ICP plasma and auxiliary gas flow rates were set at 14 L/min and

1.2 L/min respectively. Flow injection analysis and sample loading onto the HPLC column (Luna

C18 5µm, 1 mm x 250 mm, Phenomenex, Torrance, CA) were conducted using a 20 µL injection loop.

The plasma reaction products were sampled from the central channel of the ICP by a water- cooled flat aluminum sampler bearing a 2.5 mm orifice. The sampler was located 9 mm downstream of the load coil. A 4.8 mm i.d. x 100 mm length stainless steel tube served as the afterglow ionization reaction chamber.

45

4.8 mm i.d. x 100 mm metal capped glass ion length SS afterglow transfer capillary 1.5 mm reaction tube injector auxiliary MS gas endplate plasma RF load coil gas

9 mm C18 gap Column nebulizer argon plasma vacuum spray pumps chamber

injector with solvent 5.0 L/min N 20 µL loop ionization 2 drain counter-flow gas reagent grounded Al sampler with a 2.5 mm orifice LC pump 50 µL/min

Figure 2.1. Schematic diagram of the ICP-based PARCI with dual cyclonic spray chambers and atmospheric pressure ionization interface coupled to an atmospheric sampling mass spectrometer. Note that the schematic is not to scale.

46

For mass spectrometric detection, the electrospray ion source of a single quadrupole MS

(SQ300, PerkinElmer Inc., Waltham, MA) was removed and the downstream end of the reaction tube was placed in line with the inlet of the MS, 9 mm away from the endplate. A 5.00 L/min nitrogen counter flow gas sheathing over the inlet of the MS assisted in desolvation of the ions and protected the MS inlet from high temperatures of the plasma afterglow. Negative ions were drawn into the ion sampling capillary of the MS by applying a potential of 200 V to the endplate, and 600

V to the entrance of the glass capillary. An ion capillary exit potential of -150 V was used in conjunction with -30 V skimmer voltage to decluster the ions and release Cl- for detection by the

MS. The nozzle-skimmer declustering also eliminated any potential polyatomic isobaric

- interferences such as H2O∙OH . For experiments involving methanol as solvent, we intentionally reduced the ion flux by setting the Q1 offset parameter to zero (normal operation at +2 V) in single ion monitoring (SIM) mode to minimize the aging of the detector due to high background counts of chlorine. Note that this setting causes ion count reduction by a factor of 4.8 but does not impact detection limits because both signal and background are reduced by the same ratio. The 35Cl background intensity was still 15000-20000 counts/s in these conditions, ensuring that the precision of the background intensity measurements was determined by the ion source stability rather than counting statistics. Background spectra were collected by scanning m/z range of 10 to

100 with a pulse counting time of 1 ms and 10 points per mass in the absence of analyte.

Optimization and ionization studies were conducted using an integration time of 500 ms per ion in

SIM mode.

Analyte solutions of chloramphenicol and sodium chloride (both purchased from Sigma-

Aldrich, St. Louis, MO) were prepared by dissolving in LC-MS grade methanol (Chromasolv,

47

Sigma-Aldrich, St. Louis, MO) at chlorine concentrations of ~40-80 mM, then diluted to final concentration in LC-MS grade water (Chromasolv, Sigma-Aldrich, St. Louis, MO). Ionization reagent solutions and spiked chloramphenicol solutions were prepared using ACS grade sodium acetate (Fisher, Waltham, MA), lithium acetate (GFS Chemicals, Columbus, OH), potassium acetate (EM Science, Gibbstown, NJ) and LC-MS grade water, methanol, and acetonitrile. LC-MS grade methanol and water were used for HPLC gradients. All chromatograms were extracted as

.csv files and integrated using OpenChrom software [23] by manually selecting start and end points of a peak to define a baseline. The integrated peak areas were divided by a factor of 10 to report values in the unit of ion counts because of the 0.1 second time unit used in the OpenChrom integration algorithm.

2.3. Results and Discussion

2.3.1. Effect of Carrier Gas Flow Rate

The atmospheric pressure PARCI-MS interface shown in Figure 2.1 is designed to enhance electron capture and ion-molecule reactions for improved negative ion formation. Unlike the conventional ICP-MS interface where plasma undergoes free expansion into vacuum, the plasma sampling in Figure 2.1 occurs at nearly atmospheric pressure on both sides of the plasma sampling orifice. Therefore, the carrier gas flow rate supplied to the injector of the ICP torch plays a critical role in pushing the plasma reaction products into the afterglow reaction tube. Figure 2.2a demonstrates the impact of total carrier gas flow rate on sensitivity for Cl represented by peak areas from flow injections of 50 µM sodium chloride in water. In these experiments, the analyte nebulizer gas flow rate is held at 1.3 L/min to ensure constant nebulization and analyte aerosol

48 transport efficiency while makeup flow rates of 0-1.3 L/min are added via the second nebulizer and spray chamber.

The red line in Figure 2.2a demonstrates Cl sensitivity in the absence of any reagent solution sprayed through the second nebulizer/spray chamber. It is evident that the optimum sensitivity is observed at a total carrier gas flow rate of 2.2 L/min, far greater than the typical flow rates (0.8-1.2 L/min) used in conventional ICP-MS. In order to gain further insight into the effect of carrier gas flow rate on sensitivity, we measured the gas temperature at the end of the stainless- steel afterglow reaction tube using a J type thermocouple at various gas flow conditions. The results shown in Figure 2.2b (black line) demonstrate a correlation between the temperature at the end of the reaction tube and the Cl sensitivity. Increasing the total carrier gas flow rate from 1.3

L/min to 2.2 L/min increased both the temperature and the sensitivity for Cl. This could be attributed to penetration of the gas from the plasma central channel through the sampling orifice, carrying analyte plasma reaction products for ionization within the tube. Lower temperatures at the reaction tube end at gas flow rates above 2.2 L/min suggest excessive plasma cooling in these conditions. A loss in sensitivity is also observed at high carrier gas flow rates, likely as a result of compromised atomization at cooler central channel temperatures and lower electron number densities within the ionization tube.

49

Figure 2.2. Effect of total carrier gas flow rate on a) chlorine sensitivity and b) reaction tube end temperature. Peak areas for triplicate flow injections of a 50 µM chlorine sample are used as a measure of sensitivity. Each error bar represents one standard deviation from the mean.

Experimental parameters: analyte nebulizer gas flow rate = 1.3 L/min, reagent nebulizer gas flow rate = 0-1.3 L/min, RF power = 1100 W.

50

2.3.2. Effect of Ionization Reagent

A mechanistically important observation is also highlighted in Figure 2.2a for flow injections of 25 µM chloramphenicol in water (equivalent to 50 µM sodium chloride in chlorine concentration). Notably, injections of the organochlorine analyte (blue line) yielded significantly lower Cl- ion intensities compared to those of sodium chloride (red line). We hypothesized that the difference was related to presence of sodium. To test this hypothesis, sodium was added to chloramphenicol samples in the form of 50 µM sodium acetate. The green line in Figure 2.2a depicts that addition of sodium to the organochlorine analyte solution results in significant improvement of Cl- intensity. Moreover, with sodium addition, the dependence of sensitivity on carrier gas flow rate for the organochlorine analyte was similar to that recorded using sodium chloride. These results indicate an important role for sodium in Cl- ion formation.

One mechanism for the role of sodium as an easily ionizable element (EIE) is its effects on plasma properties such as electron number density [24]. Operation at high nebulizer gas flow rates leads to a cooler central channel relative to conventional ICP-MS, lowering the plasma charge density. Moreover, plasma cooling is intentionally induced in the afterglow to allow efficient formation of negative ions, resulting in further reduction of electron number density by ion- electron recombination. Addition of EIEs to the plasma can provide a new source of electrons, increasing electron number density at lower temperatures and enhancing negative ion formation.

The results outlined below discuss additional important factors for ionization in PARCI.

2.3.2.1. Independent Introduction of Ionization Reagent

51

Figure 2.3. Effect of ionization reagent (200 µM sodium acetate) solution flow rate and solvent type on chlorine sensitivity. Peak areas for triplicate flow injections of 25 µM chloramphenicol (50 µM chlorine) are used as a measure of sensitivity. Each error bar represents one standard deviation from the mean. The inset provides a zoomed-in view for the solvents with low sensitivity. Experimental parameters: analyte nebulizer gas flow rate = 1.3 L/min, reagent nebulizer gas flow rate = 0.9 L/min, RF power = 1100 W.

52

To gain further insights into ionization in PARCI and to improve sensitivity for Cl detection, we utilized the second nebulizer/spray chamber to produce sodium-containing aerosol for subsequent merging with the analyte aerosol (See Figure 2.1). We refer to the solution used for the second spray as the ionization reagent solution. The dual spray chamber configuration allows us to experiment with the solvent type and solution flow rate for introduction of reagent aerosols without affecting analyte aerosol generation and transport.

Figure 2.3 shows the effect of reagent solution flow rate and solvent type on Cl sensitivity indicated by peak areas for flow injections of 25 µM chloramphenicol in water. Chloramphenicol was selected for these studies as a representative of organochlorines and to ensure that the sodium introduced into the plasma originated mainly from the ionization reagent. A reagent solution of

200 µM sodium acetate was used for these experiments with 0.9 L/min reagent nebulizer gas flow rate and 1.3 L/min analyte nebulizer gas flow rate, providing the total optimum carrier gas flow of

2.2 L/min based on the results in Figure 2.2.

It is evident from Figure 2.3 that ionization reagent solvent type and flow rate significantly impact chlorine sensitivity. An improvement in Cl sensitivity is observed with increased reagent solution flow rate for water, methanol, and water:methanol solvents, reaching a plateau at higher flow rates. In contrast, the sensitivity decreases with increased flow rate for reagent solvents containing acetonitrile. The positive correlation of sensitivity with reagent solution flow rate using water and methanol solvents is attributed to increased sodium concentrations in the plasma and enhanced ionization. Observation of a plateau for these solvents can have multiple origins. One factor is the reduced reagent aerosol transport efficiency through the second spray chamber at higher solvent flow rates [11, 25]. A second factor may arise from ionization kinetics; that is an

53 increase in initial charge density at higher sodium concentrations (supplied by higher reagent flow rates) may be offset by faster ion-ion recombination within the denser charged cloud traveling down the reaction tube. A similar plateau for sensitivity caused by ionization kinetics has been reported for dopant flow rate in atmospheric pressure photoionization where a bipolar ion cloud induced by photoionization of the dopant is used to ionize analytes [26].

The effect of ionization reagent solvent type is more complex. Addition of organic solvent to the ionization reagent is expected to improve aerosol transport efficiency, leading to larger concentrations of sodium in the plasma. However, this effect is unlikely to be a major factor for trends in Figure 2.3 because: 1) a drastic improvement in Cl sensitivity (factors of larger than 20) is observed for methanol over water across a wide range of solvent flow rates while only improvement factors of ~5 in aerosol transport efficiency are reported for methanol compared to water [27], 2) our data presented later in section 3.5 demonstrate that only a factor of 2 improvement in sensitivity is observed between water and 80% methanol in water at an analyte solution flow rate of 50 µL/min, and 3) the sensitivity in Figure 2.3 decreases with use of acetonitrile while aerosol transport efficiency is expected to increase. These discrepancies indicate that the reagent solvent plays a critical role beyond improvement of aerosol transport through the reagent spray chamber.

Considering the drastic impact of methanolic sodium acetate on Cl sensitivity shown in

Figure 2.3, we investigated the relative contributions of EIEs and methanol by varying the concentration and type of EIEs. The results summarized in Table 2.1 indicate that sodium concentrations of up to 200 µM in methanol increase the sensitivity by a factor of 1.7 compared to neat methanol. Higher concentrations do not appear to offer further enhancement in Cl sensitivity.

54

Moreover, among Li, Na, and K introduced at 200 µM, sodium shows the largest impact followed by lithium. This trend does not follow the order of ionization potential for EIEs, suggesting that the EIE role is more complex than just providing electrons for ionization in negative mode.

Table 2.1. Effects of easily ionizable elements added to methanol on chlorine sensitivity1 Added Concentration Added Salt Normalized Peak Area2 (µM) None (neat 0 1 ± 0.053 methanol) C H O Na 50 1.24 ± 0.040 2 3 2 C2H3O2Na 200 1.68 ± 0.055 C2H3O2Li 200 1.18 ± 0.035 C2H3O2K 200 0.99 ± 0.035 C2H3O2Na 400 1.71 ± 0.024 C2H3O2Na 1000 1.42 ± 0.063 ± refers to one standard deviation 1 Ionization reagent flow rate of 30 µL/min and flow injection of 25 µM chloramphenicol were used for these experiments. 2 Peak areas are normalized to neat methanol

The results in Table 2.1 also imply that the drastic effect of methanolic sodium acetate in

Figure 2.3 is largely related to methanol rather than sodium acetate. However, one must note that significant effects (factor of 1.7) are still observed for sodium at low concentrations (< 200 µM equivalent to 5.8 ppm in methanol). Considering that sodium is always present, even in neat conditions, decoupling of methanol effect from that of sodium is not possible in our experiments.

In fact, we regularly observe 3-5 million counts/s of sodium in positive mode while operating with neat solvents. Collectively, these results suggest a complex and potentially cooperative mechanism for enhancement of Cl sensitivity by methanol and EIEs. This observation implies that chemical

55 ionization effects beyond electron capture by Cl atoms play significant roles in formation of Cl- ions.

2.3.3. Ionization Mechanism

In an attempt to decipher chemical ionization effects, we examined the background mass spectra in each condition in Figure 2.3. Major ions were identified at m/z values of 26, 42, 45, and

- - - - 46 corresponding to CN , OCN , CHO2 , and NO2 . Note that these structures are reported in the most thermodynamically stable configuration because the ions are thermalized to room

- - - temperature before sampling by the MS. We have reported detection of CN , OCN , and NO2 in low pressure PARCI with a microwave induced plasma [19, 20]. These ions result from in-plasma formation of radicals followed by electron capture in the afterglow. The sources for nitrogen, carbon, and oxygen atoms include solvents and ambient air diffusion into the plasma. The formate

- ion (CHO2 ) is observed in the presence of methanol but not acetonitrile, suggesting that the nitrogen-containing ions are thermodynamically favored when a substantial amount of nitrogen is present in the plasma as a result of acetonitrile introduction.

Figure 2.4 illustrates correlations between major background ion intensities and chlorine sensitivity based on 25 µM chloramphenicol injections. A number of trends are evident in Figure

- 2.4: 1) Both low and high Cl sensitivities are observed in Figure 2.4a at high NO2 ion intensities,

- - implying that NO2 ion is not a major factor in Cl formation mechanism. 2) A weak positive correlation is observed between formate ion intensity and Cl sensitivity for methanol-containing solvents as shown in Figure 2.4b. However, Cl sensitivity reaches a plateau at higher formate ion intensities. Accordingly, we infer that the formate ion is unlikely to be a reagent ion for Cl- formation. 3) Introduction of acetonitrile leads to significant increase in CN- and OCN- ion

56 intensities while Cl sensitivity is reduced (Figures 2.4c and 2.4d). This is consistent with high electron affinity of CN and OCN radicals (3.86 eV and 3.61 eV, respectively) [28] which makes them efficient competitors for electrons in the afterglow. Notably, both CN- and OCN- ions have higher proton affinities than Cl-. Therefore, the reduced Cl sensitivity at high intensities for CN- or OCN- implies that formation of HCl and subsequent proton transfer to these ions is not a major pathway for Cl- ion formation. In summary, investigation of background ion intensities indicates a competition for electrons in the afterglow, explaining the negative effect of acetonitrile, but the observations do not provide evidence for the role of background ions in the formation of Cl- via proton or electron transfer reactions.

It is of note that the detected background ions are the most stable products of the ionization reactions within the tube. In other words, any short-lived species that do not survive to the end of the tube would not be detected by the MS. Therefore, one explanation for the positive effect of methanol on ionization is formation of short-lived species as intermediates for ionization. For example, formation of a C-Cl bond in the afterglow followed by dissociative electron capture would enhance the intensity of Cl- ions and would be consistent with the observations above.

Further investigations of chemical ionization effects will be needed to decipher the details of ionization in PARCI.

57

Figure 2.4. Correlation plots for major background ions and chlorine sensitivity obtained at various reagent solution flow rates and solvent types with experimental conditions of Figure

2.3. Peak areas for triplicate flow injections of 25 µM chloramphenicol (50 µM chlorine) are used as a measure of sensitivity. Data points are coloured according to solvents similar to Figure 2.3.

Each error bar represents one standard deviation from the mean. Cl sensitivity is not correlated

- with NO2 ion in panel a while a loose positive correlation is observed between sensitivity and

- - HCO2 in panel b. Panels c and d show that sensitivity is reduced when high intensities of CN and

OCN- are detected.

58

Figure 2.5. Effect of plasma RF power on chlorine sensitivity and compound-independent response factors. Peak areas for triplicate flow injections of aqueous chlorine samples are used as a measure of sensitivity while the ratio of the peak areas for NaCl and chloramphenicol show compound-independent response. Each error bar represents one standard deviation from the mean.

Experimental parameters: analyte nebulizer gas flow rate = 1.3 L/min, reagent nebulizer gas flow rate = 0.9 L/min, reagent solution flow rate = 30 µL/min, reagent solution = 200 µM sodium acetate in methanol.

59

2.3.4. Effect of Plasma RF Power

The RF power applied to the plasma has a significant impact on the temperatures within the plasma and in the afterglow reaction tube. Variations in temperature could alter atomization within the plasma as well as the ionization reactions within the tube. Figure 2.5 depicts the effect of RF power on Cl sensitivity and compound-independent response factors. Peak areas from flow injections of chloramphenicol in water are used as a measure of sensitivity while the peak area ratios for injections of chloramphenicol and sodium chloride with identical chlorine concentrations

(50 µM Cl) reflect the compound-independent behavior. A peak area ratio of unity is expected for an ideal technique with compound-independent response. The two compounds are selected because they have drastically different properties, such as vaporization and decomposition temperatures and ability to undergo dissociative electron capture, making them suitable for testing compound-independent repose.

Figure 2.5 illustrates that chlorine sensitivity decreases with increasing RF power, in contrast to observations in conventional ICP-MS [7]. We attribute this trend to reduced transmission of analyte through the sampling orifice into the reaction tube. Higher RF powers lead to higher temperatures at the sampling position, reducing the density of the carrier gas entering the tube. As a result, a smaller fraction of the total sample passes through the orifice. Figure 2.5 also shows that the two compounds provide similar response factors at RF power levels ≥ 900 W.

However, chloramphenicol provides a higher response factor compared to sodium chloride at 700

W. This can be explained by a reduction in plasma temperature to a level where complete atomization is not achieved. The organochlorine compound requires lower temperatures for its

60 vaporization and breakdown and undergoes dissociative electron capture, leading to enhanced response relative to that of sodium chloride at low power levels.

2.3.5. Effect of Analyte Solvent Composition

Most applications of environmental and biological significance utilize reverse phase chromatography with solvent gradients from aqueous to high organic content. Accordingly, we have investigated the impact of analyte solvent composition on sensitivity and compound- independent response factors. Our investigations focused on water/methanol gradient because of the suppression effect observed with acetonitrile.

Figure 2.6 illustrates that Cl sensitivity increases with higher methanol fraction in the solvent. In the previous section, we established that methanol assists in ionization. However, this effect is unlikely to be the origin of the trend in Figure 2.6 because sufficient amount of methanol is already supplied by ionization reagent (30 µL/min of 200 µM sodium acetate in methanol).

Increased amounts of methanol do not lead to higher sensitivities based on the data presented in

Figure 2.3. Therefore, we conclude that the increase in the sensitivity with the methanol fraction in the analyte solvent is likely a result of improved nebulization and analyte aerosol transport efficiency through the spray chamber with organic solvents [27]. Moreover, Figure 2.6 shows that the response ratio of sodium chloride to chloramphenicol does not change as a function of analyte solvent composition, suggesting that the higher methanol load into the plasma has minimal impact on atomization efficiency at the solvent flow rates studied in this report.

61

Figure 2.6. Effect of methanol content in analyte solvent on chlorine sensitivity and compound-independent response factors. Peak areas for triplicate flow injections of aqueous chlorine samples are used as a measure of sensitivity while the ratio of the peak areas for NaCl and chloramphenicol show compound-independent response. Peak areas are normalized to that of analyte in water. Each error bar represents one standard deviation from the mean. Experimental parameters: analyte nebulizer gas flow rate = 1.3 L/min, reagent nebulizer gas flow rate = 0.9

L/min, reagent solution flow rate = 30 µL/min, reagent solution = 200 µM sodium acetate in methanol, RF power = 1100 W.

62

Notably, the positive effect of methanol on Cl sensitivity shown in Figures 2.3 and 2.6 is in contrast to the ion suppression observed in ICP-MS with organic solvents for this element [29].

While ICP-MS heavily relies on high plasma gas temperature for ion formation, PARCI takes advantage of ionization at low temperatures. Therefore, ionization in PARCI is less susceptible to cooling effects of organic solvents, particularly for high ionization potential elements.

2.3.6. Analytical Figures of Merit

The analytical performance of PARCI was examined at optimized conditions (200 µM methanolic sodium acetate as the ionization reagent infused at 30 µL/min, 0.9 L/min ionization reagent nebulizer gas flow, 1.3 L/min analyte nebulizer flow, 1100 W RF power) in two modes:

1) direct infusion and 2) LC mode. The direct infusion was achieved by a 60 µL injection of 5 µM chloramphenicol (10 µM Cl) in water at a flow rate of 50 µL/min, providing a flat top peak.

Notably, the ion flux reduction was not applied for this experiment to investigate the potential of

PARCI in terms of sensitivity. Table 2.2 shows the ion intensities for 35Cl- and 37Cl- ions at the baseline and the flat top portion of the injection. The sensitivities and detection limits based on these results are summarized in Table 2.3 and are compared to those of ICP-MS/MS and ICP-high resolution MS reported in the literature.

Table 2.2. Chlorine ion intensities from flat-top peak flow injection of chloramphenicol in water Injected Cl Average 35Cl Average 37Cl 35Cl/37Cl Isotope Ratio Concentration (µM) Intensity (cps) Intensity (cps)

0 84629 ± 1116 26086 ± 568 3.24 ± 0.08 10 255252 ± 7348 80232 ± 2376 3.1 ± 0.2 ± refers to one standard deviation

63

The results in Table 2.3 indicate that PARCI-MS provides at least two-fold larger detected

Cl ion intensity (cps per ng/mL) compared to both ICP-MS/MS and ICP-high resolution MS. This finding is even more notable considering that the MS used in our experiments is designed for molecular mass spectrometry and is not optimized for ion transmission at low mass. Moreover, the sensitivity for ICP-MS/MS in the last row of Table 2.3 is reported for methanol as analyte solvent.

Our results above show that an additional factor of two improvement in PARCI’s sensitivity is achieved by using methanol as analyte solvent, further highlighting the potential of PARCI for high sensitivity analysis of Cl. The detection limits, however, are slightly higher for PARCI mainly because of large Cl background in our experiments evident from background equivalent concentration (BEC) of 170 ng/ml. It is of note that the high background in our experiments appears to be Cl contamination rather than isobaric interferences evident from similar 35Cl/37Cl isotope ratio values for analyte and the background listed in Table 2.2. Reducing the contamination and improvements in ionization reagent introduction setup (e.g. concentric versus cross flow aerosol mixing) are expected to further improve the detection limits for PARCI.

For LC mode, chromatography was performed with a gradient of methanol in water starting at 5% and ramping to 80% in 8 minutes, holding at 80% for 2 minutes, followed by ramping down to 5% over 4 minutes. A 12-cm long 100 µm i.d. fused silica capillary inserted into the solution uptake barrel of the HEN was used to deliver the LC eluent into the nebulizer with minimal post- column dead volume. Figure 2.7 shows chromatographic peak shape and linearity of chlorine response for LC-PARCI-MS of chloramphenicol. Full width at half maximum for the chloramphenicol peak was 0.15 min while the asymmetry (tailing) factor measured at 5% of the maximum height was 1.5. The asymmetry factor is below the threshold of 2.0 established by US

64

FDA for validated methods, demonstrating minimal distortion of peaks by the detector. Better peak shapes may be obtained by optimization of the chromatographic method.

Table 2.3. Comparison of chlorine detection limits and sensitivities Instrument Cl LOD Sensitivity BEC 35Cl/37Cl Isotope (ng/mL) (cps per (ng/mL) Isotope ng/mL) ratio

PARCI-MS1 35 6.97 481 176 3.1* 37 11.18 153 171

HR-ICP-MS2 35 7.01 222 0.17 N/A** resolution = 4000 37 N/A** N/A** N/A**

HR-ICP-MS2 35 3.25 34 0.08 2.6* resolution = 37 4.18 13 0.08 10,000

ICP-MS/MS3 35 0.28 176 1.83 N/A

ICP-MS/MS4 35 1 287 52 N/A

* calculated as sensitivity of 35Cl divided by sensitivity of 37Cl ** values not reported because resolution was not sufficient for baseline separation of isobaric interferents 1 This work based on data in Table 2.2 2 Reference [7] 3 Reference [32] 4 Reference [5], Note: methanol was used as solvent for direct infusion of analyte

Nearly two orders of magnitude of linear response is also shown in Figure 2.7b for chloramphenicol injected at concentrations of 0.05-2.5 µM (0.1-5 µM chlorine) with R2 = 0.9996.

Calibration points were also obtained at higher concentrations of 10, 20, and 50 µM chlorine, however departure from linearity was observed at these concentrations as shown in the inset of

Figure 2.7b. Deviations from linearity could arise from ionization, ion transfer into the mass spectrometer, and ion transmission effects within the mass spectrometer. The exact contributions of these factors remain to be investigated. 65

Figure 2.7. a) Gradient chromatographic elution of 2.5 µM chloramphenicol injected onto a

C18 column and detected by PARCI-MS and b) calibration curve for chlorine (0.1-5 µM) based on peak areas from the chromatograms. An extended calibration curve depicting departure from lienarity is shown as inset. Experimental parameters: analyte nebulizer gas flow rate = 1.3 L/min, reagent nebulizer gas flow rate = 0.9 L/min, reagent solution flow rate = 30

µL/min, reagent solution = 200 µM sodium acetate in methanol, RF power = 1100 W. 66

The limit of detection (LOD) for chlorine using LC-PARCI-MS was determined to be

0.029 µM (1 ng/mL, 21 pg on column, 2.3 pg/s) based on three times the standard deviation of the baseline prior to the peak obtained for 0.1 µM Cl injection. This corresponds to an LOQ of 0.097

µM (3.3 ng/mL, 69 pg on column). These values compare favorably to recent investigations using

HPLC-ICP-MS/MS with hydrogen as the reaction gas. For example, an LOQ of 50 ng/mL (2.5 ng on column) is reported for peaks eluted with 85% methanol [5]. However, it must be noted that peak-to-peak noise as opposed to standard deviation was used in this study to calculate LOQ.

Based on the definition used in our investigations, an LOQ of 500 pg on column is estimated for

HPLC-ICP-MS/MS, about 7 times larger than LOQ reported in our experiments. In other studies, a chlorine LOD of 0.29 µM (10.3 ng/mL, 206 pg on column) has also been reported for HPLC-

ICP-MS/MS with isocratic elution using 20% acetonitrile [30], while orders of magnitude inferior

LOD (128 µM) is obtained for single quadrupole LC-ICP-MS using water as solvent and a high temperature column for separation [31]. In summary, LC-PARCI-MS currently offers favorable analytical performance for Cl detection, but further improvements in sensitivity are expected with optimization of ionization chemistry and interface design.

2.4. Conclusions

The studies above demonstrate that an atmospheric pressure PARCI source comprised of an Ar-ICP coupled to a post-plasma reaction tube can effectively generate negative ions for compound-independent elemental and isotopic analysis of liquid samples. In particular, we have established that this ionization strategy offers chlorine detection limits comparable to ICP-MS/MS and ICP-high resolution MS but using a single quadrupole MS designed for molecular mass spectrometry with no modifications to the MS.

67

Unlike ICP-MS in which ion sampling is conducted with minimal ion-molecule reactions,

PARCI relies on post-plasma neutral-electron and ion-neutral reactions for ion formation.

Ionization at low temperatures and low electron number densities within the reaction tube in

PARCI requires additional electron donors for efficient negative mode ionization. In our previous investigations using low-pressure helium plasma, the electron donor was the nitrogen gas introduced in the afterglow, leading to Penning ionization by helium metastable atoms. In the current study, however, the electrons are provided by introducing a low ionization potential element (e.g. sodium) into the Ar ICP.

The drastically different ionization mechanisms between PARCI-MS and ICP-MS also lead to major differences for solvent and RF power effects. The organic solvents and lower RF powers generally deteriorate Cl sensitivity in ICP-MS. In contrast, methanol and low RF powers lead to significant improvements of sensitivity in PARCI-MS while acetonitrile reduces the sensitivity. The suppressive effects of acetonitrile appear to be related to depletion of electrons in the afterglow by CN and OCN radicals. But the mechanism for dramatic enhancement of Cl- signal by methanol remains unclear. A significant amount of formate ion is detected when methanol is introduced into the plasma, however, investigation of various ionization reagent flow rates shows that the Cl- sensitivity is only loosely correlated with formate ion intensity. Further, the sensitivity enhancement by sodium added to methanol surpasses those observed by addition of lithium and potassium. These observations indicate that additional mechanisms beside electron generation may be at play for effects of easily ionizable elements, reflecting the complexity of chemical ionization in the afterglow.

68

Our results clearly highlight the sensitivity and versatility of atmospheric pressure PARCI-

MS for Cl analysis. Further fundamental and optimization investigations are underway in our group to improve the performance of PARCI and to extend its application to other non-metals. The compatibility of atmospheric pressure PARCI with molecular MS instruments paves the way for integration of elemental methods with molecular MS toward a comprehensive analytical platform.

2.5. References

1. Niu, H., Houk, R.S.: Fundamental aspects of ion extraction in inductively coupled plasma mass spectrometry. Spectrochim. Acta Part B At. Spectrosc. 51, 779–815 (1996). https://doi.org/10.1016/0584-8547(96)01506-6

2. Evans, E.H., Giglio, J.J.: Interferences in inductively coupled plasma mass spectrometry.

A review. J. Anal. At. Spectrom. 8, 1–18 (1993). https://doi.org/10.1039/JA9930800001

3. Pröfrock, D., Leonhard, P., Wilbur, S., Prange, A.: Sensitive, simultaneous determination of P, S, Cl, Br and I containing pesticides in environmental samples by GC hyphenated with collision-cell ICP-MS. J. Anal. At. Spectrom. 19, 623–631 (2004). https://doi.org/10.1039/B310530J

4. Böting, K., Treu, S., Leonhard, P., Heiß, C., Bings, N.H.: First experimental proof of asymmetric charge transfer in ICP-MS/MS (ICP-QQQ-MS) through isotopically enriched oxygen as cell gas. J. Anal. At. Spectrom. 29, 578–582 (2014). https://doi.org/10.1039/C3JA50234A

5. Klencsár, B., Bolea-Fernandez, E., Flórez, M.R., Balcaen, L., Cuyckens, F., Lynen, F.,

Vanhaecke, F.: Determination of the total drug-related chlorine and bromine contents in human blood plasma using high performance liquid chromatography–tandem ICP-mass spectrometry

69

(HPLC–ICP-MS/MS). J. Pharm. Biomed. Anal. 124, 112–119 (2016). https://doi.org/10.1016/j.jpba.2016.02.019

6. Bolea-Fernandez, E., Balcaen, L., Resano, M., Vanhaecke, F.: Overcoming spectral overlap via inductively coupled plasma-tandem mass spectrometry (ICP-MS/MS). A tutorial review. J. Anal. At. Spectrom. (2017). https://doi.org/10.1039/C7JA00010C

7. Bu, X., Wang, T., Hall, G.: Determination of halogens in organic compounds by high resolution inductively coupled plasma mass spectrometry (HR-ICP-MS). J. Anal. At. Spectrom.

18, 1443–1451 (2003). https://doi.org/10.1039/B306570G

8. Egorova, T., Dietiker, R., Hattendorf, B., Günther, D.: Characteristics of an ion funnel as ion guide in an inductively coupled plasma mass spectrometer. Spectrochim. Acta Part B At.

Spectrosc. 76, 40–47 (2012). https://doi.org/10.1016/j.sab.2012.06.010

9. Pack, B.W., Broekaert, J.A.C., Guzowski, J.P., Poehlman, J., Hieftje, G.M.: Determination of Halogenated Hydrocarbons by Helium Microwave Plasma Torch Time-of-Flight Mass

Spectrometry Coupled to Gas Chromatography. Anal. Chem. 70, 3957–3963 (1998). https://doi.org/10.1021/ac970534r

10. Jorabchi, K., Montaser, A.: New coil geometries for the study of secondary discharge in atmospheric-pressure helium inductively coupled plasma mass spectrometry. Spectrochim. Acta

Part B At. Spectrosc. 59, 1471–1479 (2004). https://doi.org/10.1016/j.sab.2004.07.009

11. Jorabchi, K., McCormick, R., Levine, J.A., Liu, H., Nam, S.-H., Montaser, A.: High efficiency nebulization for helium inductively coupled plasma mass spectrometry. Spectrochim.

Acta Part B At. Spectrosc. 61, 945–950 (2006). https://doi.org/10.1016/j.sab.2006.08.003

70

12. Heitkemper, D., Creed, J., Caruso, J.A.: Helium Microwave-Induced Plasma Mass

Spectrometric Detection for Reversed-Phase High-Performance Liquid Chromatography. J.

Chromatogr. Sci. 28, 175–181 (1990). https://doi.org/10.1093/chromsci/28.4.175

13. Chtaib, M., Schmit, J.-P.: Negative-ion mode in inductively coupled plasma mass spectrometry. J. Anal. At. Spectrom. 3, 315–318 (1988). https://doi.org/10.1039/JA9880300315

14. Vickers, G.H., Wilson, D.A., Hieftje, G.M.: Detection of negative ions by inductively coupled plasma mass spectrometry. Anal. Chem. 60, 1808–1812 (1988). https://doi.org/10.1021/ac00168a031

15. Fulford, J.E., Quan, E.S.K.: Negative Ions in Inductively Coupled Plasma Mass

Spectrometry. Appl. Spectrosc. 42, 425–428 (1988)

16. Pupyshev, A.A., Surikov, V.T.: Application of negative ions in inductively coupled plasma-mass spectrometry. Spectrochim. Acta Part B At. Spectrosc. 59, 1021–1031 (2004). https://doi.org/10.1016/j.sab.2004.05.011

17. Wang, H., Lin, N., Kahen, K., Badiei, H., Jorabchi, K.: Plasma-Assisted Reaction

Chemical Ionization for Elemental Mass Spectrometry of Organohalogens. J. Am. Soc. Mass

Spectrom. 25, 692–695 (2014). https://doi.org/10.1007/s13361-013-0816-5

18. Lin, N., Wang, H., Kahen, K., Badiei, H., Jorabchi, K.: Gas Chromatography Plasma-

Assisted Reaction Chemical Ionization Mass Spectrometry for Quantitative Detection of Bromine in Organic Compounds. Anal. Chem. 86, 7954–7961 (2014). https://doi.org/10.1021/ac501964u

19. Wang, H., Minardi, C.S., Badiei, H., Kahen, K., Jorabchi, K.: High-sensitivity elemental ionization for quantitative detection of halogenated compounds. Analyst. 140, 8177–8185 (2015). https://doi.org/10.1039/C5AN01958C

71

20. Haferl, P.J., Zheng, K., Wang, H., Jorabchi, K.: Elemental quantitation of carbon via production of polyatomic anions in gas chromatography-plasma assisted reaction chemical ionization mass spectrometry. Anal. Bioanal. Chem. 409, 3843–3851 (2017). https://doi.org/10.1007/s00216-017-0328-4

21. Pröfrock, D., Prange, A.: Inductively Coupled Plasma–Mass Spectrometry (ICP-MS) for

Quantitative Analysis in Environmental and Life Sciences: A Review of Challenges, Solutions, and Trends. Appl. Spectrosc. 66, 843–868 (2012). https://doi.org/10.1366/12-06681

22. Wolf, K.D., Balcaen, L., Walle, E.V.D., Cuyckens, F., Vanhaecke, F.: A comparison between HPLC-dynamic reaction cell-ICP-MS and HPLC-sector field-ICP-MS for the detection of glutathione-trapped reactive drug metabolites using clozapine as a model compound. J. Anal.

At. Spectrom. 25, 419–425 (2010). https://doi.org/10.1039/B921638C

23. Wenig, P., Odermatt, J.: OpenChrom: a cross-platform open source software for the mass spectrometric analysis of chromatographic data. BMC . 11, 405 (2010). https://doi.org/10.1186/1471-2105-11-405

24. Luis Todolí, J., Gras, L., Hernandis, V., Mora, J.: Elemental matrix effects in ICP-AES. J.

Anal. At. Spectrom. 17, 142–169 (2002). https://doi.org/10.1039/B009570M

25. Todolí, J.-L., Hernandis, V., Canals, A., Mermet, J.-M.: Comparison of characteristics and limits of detection of pneumatic micronebulizers and a conventional nebulizer operating at low uptake rates in ICP-AES. J. Anal. At. Spectrom. 14, 1289–1295 (1999). https://doi.org/10.1039/A900598F

26. Robb, D.B., Blades, M.W.: Effects of Solvent Flow, Dopant Flow, and Lamp Current on

Dopant-Assisted Atmospheric Pressure Photoionization (DA-APPI) for LC-MS. Ionizationvia

72

Proton Transfer. J. Am. Soc. Mass Spectrom. 16, 1275–1290 (2005). https://doi.org/10.1016/j.jasms.2005.03.017

27. Todoli, J.L., Canals, A., Hernandis, V.: Behaviour of a single-bore high-pressure pneumatic nebulizer operating with alcohols in inductively coupled plasma atomic emission spectrometry. J. Anal. At. Spectrom. 11, 949–956 (1996). https://doi.org/10.1039/JA9961100949

28. Bradforth, S.E., Kim, E.H., Arnold, D.W., Neumark, D.M.: Photoelectron spectroscopy of

CN−, NCO−, and NCS−. J. Chem. Phys. 98, 800–810 (1993). https://doi.org/10.1063/1.464244

29. Jensen, B.P., Gammelgaard, B., Hansen, S.H., Andersen, J.V.: Comparison of direct injection nebulizer and desolvating microconcentric nebulizer for analysis of chlorine-, bromine- and iodine-containing compounds by reversed phase HPLC with ICP-MS detection. J. Anal. At.

Spectrom. 18, 891–896 (2003). https://doi.org/10.1039/B304651F

30. Sugiyama, N.: Analysis of sulfur, phosphorus, silicon and chlorine in N-methyl-2- pyrrolidone. In: AGILENT 8800 ICP-QQQ APPLICATION HANDBOOK. pp. 16–18. Agilent

Technologies (2015)

31. de Vlieger, J.S.B., Giezen, M.J.N., Falck, D., Tump, C., van Heuveln, F., Giera, M., Kool,

J., Lingeman, H., Wieling, J., Honing, M., Irth, H., Niessen, W.M.A.: High temperature liquid chromatography hyphenated with ESI-MS and ICP-MS detection for the structural characterization and quantification of halogen containing drug metabolites. Anal. Chim. Acta. 698,

69–76 (2011). https://doi.org/10.1016/j.aca.2011.04.053

32. Nakano, K.: Ultra-low level determination of phosphorus, sulfur, silicon and chlorine using the Agilent 8900 ICP-QQQ. In: Ultra-low level determination of phosphorus, sulfur, silicon and chlorine using the Agilent 8900 ICP-QQQ. pp. 1–5. Agilent Technologies (2016)

73

Chapter III

Mechanistic Insights into Chloride Ion Detection from the Atmospheric-Pressure

Afterglow of an Argon Inductively Coupled Plasma†

3.1. Introduction

Inductively coupled plasma mass spectrometry (ICP-MS) is widely used for high-sensitivity elemental analysis. However, the capabilities of this technique are limited where analysis of low- mass and high-ionization-potential elements is needed. In particular, chlorine detection is hindered by low efficiency of Cl+ formation in the argon plasma [1] and by polyatomic isobaric interferences such as 16O18OH+. Importantly, organochlorines have emerged as a prominent class of compounds with environmental and biological significance. For example, 23 of 28 chemicals adopted by the

Stockholm Convention for global regulation are chlorinated [2]. Our examination of publicly available databases show that about 13% of the nearly 8000 drugs in the DrugBank [3] database contain chlorine, and 46% of the 768 compounds measured by the FDA’s Pesticide Residue

Monitoring Program [4] are chlorinated. Further, characterization of chlorine-containing water disinfection byproducts continues to be an active area of investigation [5–7]. Therefore, high- sensitivity elemental detection and isotopic analysis of chlorine is needed to help elucidate the environmental and biological impact of these compounds and their transformation products.

† This chapter has been modified, integrated with supplementary information, and reprinted with permission (see Appendix) from Lesniewski, J.E., McMahon, W.P., and Jorabchi, K.: Mechanistic insights into chloride ion detection from the atmospheric-pressure afterglow of an argon inductively coupled plasma. J. Anal. At. Spectrom. 33, 1981- 1992 (2018). Copyright 2018 Royal Society of Chemistry. The authors would like to acknowledge that this material is based upon work supported by the National Science Foundation (NSF) under CHE-1507304. The autors thank PerkinElmer Inc. for the loan of the single quadrupole and time of flight mass spectrometers and the ICP generator used in these studies. The authors are grateful to Dr. Hamid Badiei of PerkinElmer for constructive discussions during the course of this research. The authors would also like to disclose that Kaveh Jorabchi is an inventor of plasma assisted reaction chemical ionization (PARCI) patented under US patent #9,966,243. 74

To address challenges in chlorine elemental analysis, instrumental techniques using ICP-high- resolution MS and ICP-MS/MS have emerged in recent years, focusing on reduction of isobaric interferences [8–15]. On the other hand, we have pursued improved ionization via development of plasma-assisted reaction chemical ionization (PARCI) as a chemical strategy to enhance sensitivity and to reduce isobaric interferences [16–19]. In particular, we recently reported a PARCI source using an argon ICP for detection of Cl- ions from chlorinated compounds [19]. In the ICP-based

PARCI, analyte aerosols are introduced into an ICP similar to a conventional ICP-MS. However, the plasma products in PARCI are sampled into an atmospheric-pressure reaction tube for post- plasma ion-neutral reactions. This is in contrast to the conventional ICP-MS where ions (plasma products) are sampled directly from the plasma into the first vacuum stage of the MS, and ion- neutral reactions during ion sampling are avoided [1].

The post-plasma ion-neutral reactions in PARCI extend the ion formation pathways beyond thermal ionization in conventional ICP-MS, offering new chemical ionization strategies for elemental analysis. Further, the reaction tube situated after the plasma lowers the plasma gas temperature, enabling detection of ions using readily available atmospheric-sampling mass spectrometers (e.g. LC-MS instruments). This alleviates the need for a dedicated elemental mass analyzer and expands the elemental MS capabilities such as negative ion detection currently unavailable in commercial ICP-MS instruments.

Importantly, we have demonstrated that ICP-based PARCI-MS offers higher sensitivity for chlorine (detected ions per second per ppb of Cl) via detection of Cl- compared to that achieved via Cl+ detection in ICP-MS/MS and ICP-high resolution MS [8, 12, 19]. Notably, we reported drastic sensitivity improvement in Cl- detection using PARCI-MS upon concurrent introduction of

75 sodium and methanol into the plasma [19]. We attributed this observation to chemical ionization effects in the afterglow, however, an exact mechanism of action for methanol and sodium remained to be investigated. In the current report, we characterize the impact of plasma and sampling operating parameters on Cl- detection and offer insights into critical ion-neutral reactions in the plasma afterglow. These investigations provide insights into the effects of methanol and sodium and offer a roadmap for further improvements of PARCI by developing element-specific ionization strategies.

3.2. Experimental Details

3.2.1. Plasma Assisted Reaction Chemical Ionization (PARCI)

The ICP-based PARCI-MS setup used in this study has been described previously [19]. An updated diagram of the setup is shown in Figure 3.1 showing minor modifications to improve sensitivity and provide further control for experiments. Briefly, the analyte solution was introduced into a high efficiency nebulizer (HEN-120, Meinhard, Golden, CO) coupled to a cyclonic spray chamber (Twister Cyclonic P/N 20-809-0365, Glass Expansion, Pocasset, MA) using a 20-μL injection loop at a water flow rate of 200 μL/min supplied by an HPLC pump (SCL10AVP controller and LC10AD pump, Shimadzu, Columbia, MD). The nebulizer was operated at a constant gas flow rate of 1.25 standard liters per minute (SLPM) measured using a mass flow meter

(Model 8270, Matheson, Montgomeryville, PA). The analyte aerosol was mixed with ionization reagent aerosol (200 μM sodium acetate in methanol) emerging from a second nebulizer (HEN-

170, Meinhard, Golden, CO) and spray chamber (Twister Cyclonic P/N 20-809-0222, Glass

Expansion, Pocasset, MA) operated at nebulizer gas flow rate of 0.55 SLPM (supplied by the internal mass flow controller of the ICP generator) and solution flow rate of 30 µL/min provided

76 by a syringe pump (Model 100, KD Scientific, Holliston, MA). The mixed aerosol was directed into a tangential mixer (ML151008N, Meinhard, Golden, CO) where an argon makeup gas of varying flow rates (supplied by a mass flow controller, Model 946, MKS, Andover, MA) was added. This configuration allowed variation of total aerosol carrier gas flow rate with minimal impact on analyte and ionization reagent aerosol production and transport to the ICP. The plasma was sustained using a stand-alone ICP generator with a fully shielded torch box (NexION 2000,

PerkinElmer Inc., Waltham, MA).

To control the torch box exhaust flow rate, a 4” inline fan (Model GLFANXINLINE4, iPower

Inc., Industry, CA) was installed on the exhaust manifold of the torch box. The fan velocity was controlled by adjusting the input voltage using a variac (model Powerstat 3PN116B, ISE Inc.,

Cleveland, OH). An air velocity sensor (Model f661, Degree Controls Inc., Milford, NH) installed at the bottom air intake grid of the ICP torch box was used to monitor the exhaust flow.

An aluminum cylindrical interface provided plasma sampling for afterglow reactions. The interface provided a ¼” channel in its center terminating in a flat sampling orifice (2-mm i.d. and

1 mm thick). A stainless-steel reaction tube with a ¼” o.d. and 0.19” i.d. was inserted into the channel. Compared to the previous report [19], the reaction tube length was shortened to 50 mm and the tube was sealed to the interface using a graphite ferrule (Valco Instrument Co., Houston,

TX) to minimize air diffusion into the post-plasma reaction region. The aluminum interface was threaded onto the water-cooled aluminum plate of the ICP generator. The plate was bolted to the front of the ICP torch box and the sampling orifice was aligned with the injector tube orifice.

Operating parameters for PARCI are summarized in Table 3.1.

77

torch box Exhaust flow 1.5 mm tangential injector 4.8 mm i.d. x 50 mixer auxiliary plasma RF load mm long SS crossflow gas gas coil reaction tube mixer (tee)

nebulizer Ions to AP MS 1.25 L/min 0.55 L/min argon argon makeup gas (0-1.0 L/min) argon plasma aluminum interface air intake with 2.5 mm injector with grid sampling solvent orifice 20 µL loop drain exhaust velocity 30 µL/min sensor 200 µM LC pump sodium acetate 200 µL/min in methanol water

Figure 3.1. Schematic diagram of ICP-PARCI ion source. Note that diagram is not to scale.

78

Table 3.1. Operating parameters for ICP-based PARCI Parameter Value Plasma gas 14 L/min Auxiliary gas 1.2 L/min RF power 1100W Injector tip diameter 1.5 mm Analyte nebulizer gas 1.25 SLPM Reagent nebulizer gas 0.55 SLPM Makeup gas 0-1 SLPM (0.2 SLPM optimum conditions) Analyte solution flow rate 200 µL/min LC-MS water Reagent solution 200 µM sodium acetate in LC-MS grade methanol Reagent solution flow rate 30 µL/min Exhaust velocity 3.8, 4.8, 5.8, 6.8 m/s (5.8 m/s optimum conditions) Sampler Flat aluminum orifice, 2 mm id, 1 mm thick Sampling depth 9 mm downstream of the load coil Reaction tube 50 mm stainless steel tube, ¼” o.d., 0.19” i.d

3.2.2. Mass Spectrometers

A single quadrupole mass spectrometer (SQ 300, PerkinElmer Inc., Waltham, MA) was used for detection of ions emerging from the reaction tube. The electrospray ionization source from the instrument was removed and the PARCI source was placed such that the end of the reaction tube was aligned with the inlet of the MS and 5-10 mm away from the end plate. A nitrogen counterflow gas of 5 L/min at room temperature was used to limit the introduction of neutrals into the MS inlet. Negative ions were drawn into the inlet of the MS by applying potentials of +200 V to the end plate and +600 V to the glass capillary entrance. Background scans were collected in m/z range of 15 to 350 using 500 µs pulse counting time per point, 10 points per mass, and were averaged for 30 seconds. Both high declustering (-150 V capillary exit, -20 V skimmer) and low declustering (-30 V capillary exit, -20 V skimmer) conditions were used for background ion collections.

79

Cl sensitivity measurements were conducted in selected ion monitoring (SIM) mode at m/z values of 35 and 37 using high declustering potentials (see above) to convert all species to Cl-. An integration time of 250 ms per ion was applied. Because of the high Cl background the ion optics were detuned to reduce the ion flux and to minimize premature aging of the detector. This was effected by setting the Q1 offset parameter to zero (normally it is set to +2 V), producing a suppression factor of 4.8 for both signal and background.

For ion identity confirmations, a time-of-flight MS (Axion, PerkinElmer Inc., Waltham, MA) was utilized. Because of the detector sensitivity to argon in this instrument, the PARCI source was placed off axis and a counterflow gas of 6.5 L/min was used to minimize argon introduction into the vacuum chamber. These settings compromised the sensitivity but the exact mass from the TOF provided confidence for the identification of the ions.

3.2.3. Argon Gas Velocity and Temperature Measurements

Argon gas flow measurements in the reaction tube were based on the velocity and the temperature of the gas emerging from the tube outlet. Temperature was measured using a thermocouple and velocity was measured by tracking the movement of water droplets entrained in the center of the argon flow emerging from the reaction tube. To measure the argon gas velocity at the center of the reaction tube, we utilized a particle tracking approach. Water droplets with a diameter of ~30 μm were created on demand and were entrained in the gas flow emerging from the reaction tube. The on-demand droplet generator consisted of a 0.5 mL polypropylene water reservoir connected through a 50-cm fused silica capillary (i.d.: 100 μm, o.d.: 360 μm, Polymicro

Technologies, Phoenix, AZ, USA) to a borosilicate pulled glass tip (i.d.: 10 μm, ± 20%, WPI,

Sarasota, FL, USA). The glass tip was encased by a ceramic piezoelectric tube (PI Ceramic,

80

Lederhose, Germany) and was positioned face up and orthogonal to the reaction tube approximately 6 mm below the center of the tube. The piezoelectric actuator was driven by a trapezoidal pulse (amplitude: 55.6 V, rise time: 35 μs, duration: 80 μs, fall time: 60 μs) created by an arbitrary waveform generator (Dataman, Model: 530, Orange City, FL, USA) coupled to an amplifier (Trek, Model 2205, Lockport, New York, USA). Droplets were imaged using a triggered camera (Model DMK 42BUC03, The Imaging Source, LLC, Charlotte, NC) attached to the objective of a stereoscopic microscope.

Two green LEDs synchronized with the droplet generation were used for illumination and droplet tracking. Timing events were controlled using a microprocessor (UNO, Arduino AG,

Turin, Italy) at 20 Hz cycle time, starting with camera trigger followed by a 350 µs delay (camera ready time) before triggering the waveform generator for droplet production. A delayed pulser

(PM 5715, Philips, Eindhoven, Netherlands) simultaneously triggered with the droplet generator provided the illumination and image capture pulses to LEDs (8.2 μs duration, 4 V pulse). The droplet trajectory was tracked by varying the delay time of the LED pulser.

The droplets were ejected orthogonal to the gas flow and droplet entrainment was achieved 2 mm downstream of the tube evident from horizontal movement of the droplets with the gas flow.

Once entrained, the droplets were captured at two LED firing times. The time difference between the two firings (1-6 ms) was measured using a digital oscilloscope (DS1102E, Rigol, Beaverton,

OR). Five replicate images were captured for each LED firing time to provide an average location of the droplets for each setting. The droplet travel distance between the two LED firings was measured using Image J software package in pixels and then converted to mm (3-6 mm) based on

81 calibration using a calibrated microscope slide (model SK2, OptixCam, Roanoke, VA). The measurements provided relative standard deviations of 2-9% for the calculated velocity.

The gas temperature at the end of the reaction tube was measured using a thermocouple positioned at the center of the tube.

3.2.4. Computational Methods

Reaction thermochemistries were calculated at a theory level of ωB97xD/aug-cc-pVTZ using the approximate temperature at the end of the reaction tube in conditions where Cl sensitivity was optimized (600K). This level of theory has been successfully used to describe the thermochemical properties of halide-neutral adducts [20]. Structures were first optimized using the MMFF94 force field in the program [21] to generate initial geometries for each species. The geometries were then optimized at the final theory level using 16 [22]. In cases where more than one conformer was possible for a given ion or molecule, the lowest energy conformer was used.

3.2.5. Sample Preparation and Data Analysis

Aqueous solutions of chloramphenicol and sodium chloride with 25 μM chlorine were used in experiments. All solvents used were LC-MS grade (Chromasolv, Sigma-Aldrich, St. Louis,

MO). Chloramphenicol and sodium chloride were purchased from Sigma-Aldrich, St. Louis, MO.

Analyte solutions were prepared by dissolving in methanol to concentrations of ~40-80 mM followed by dilution in water to final concentration of 25 μM chlorine in water.

Ionization reagent solution was prepared using ACS grade sodium acetate (Fisher, Waltham,

MA) by dissolving to a concentration of ~ 96 mM in methanol, followed by dilution to 200 μM in methanol. All selected ion chromatograms from flow injections were exported as text files, and then manually integrated using the trapezoid rule by selecting the beginning and end points of a

82 baseline for each peak. The background spectra were centroided and extracted as .csv files for linear regression analyses of ion intensities.

3.3. Results and Discussion

To investigate Cl- formation mechanism in PARCI, three major processes should be considered: 1) vaporization and breakdown of chlorinated analytes in the plasma, resulting in compound-independent response factors, 2) transfer of plasma products into the reaction tube, and

3) ion-neutral reactions of plasma products in the reaction tube. These processes are influenced by plasma operating parameters and plasma modifiers (e.g. sodium and organic solvents). Among the plasma operating parameters, aerosol gas flow rate and RF power exert the largest effects on the temperature and chemical composition of the sampling position in the plasma. Another important operating parameter is the torch box exhaust flow rate which influences the pressure in the torch box. Accordingly, flow of argon from the plasma into the reaction tube is reduced at increased exhaust flow rates. Note that this is in contrast to conventional ICP-MS where the effect of exhaust flow rate on plasma sampling is minimized because of the large pressure difference on the two sides of the sampler cone.

In the following sections, we examine the influence of aerosol carrier gas flow rate and exhaust flow rate on the three processes discussed above. We have used a constant RF power of 1100 W to limit the number of variables and to allow detailed exploration of the other parameters. This power level was selected based on good analytical performance for Cl- detection in our previous investigations [19]. Moreover, all experiments utilized concurrent introduction of 200 μM methanolic sodium acetate into the plasma as ionization reagent (plasma modifier) to improve Cl-

83 sensitivity. For the evaluation of exhaust effects, the air velocity at the air intake grid of the torch box (see Experimental section) was used as an indicator for exhaust flow rate.

3.3.1. Analyte Breakdown

Uniformity of chlorine response factors (detected ions per mole of chlorine in the sample) among different compounds offers insights into analyte vaporization and breakdown efficiency.

We have selected chloramphenicol and sodium chloride as test analytes because they represent drastically different chemical properties for Cl. Figure 3.2 shows the impact of total aerosol carrier gas flow rate and exhaust velocity setting on the ratio of response factors between the two compounds. Ideally, one expects a response factor ratio of unity for compound-independent elemental ionization. Ratios of close to unity are observed at total carrier gas flow rates ≤ 2.2

SLPM. At higher flow rates, significant deviation from unity is observed, indicating extensive cooling of the central channel and compromised analyte vaporization and breakdown [1, 23].

Notably, the exhaust flow rate does not significantly influence response factor ratios in Figure

3.2 for aerosol flow rates ≤ 2.2 SLPM where compound-independent behaviors is observed. In contrast, the exhaust setting drastically impacts the reaction tube temperature as discussed in later sections. Accordingly, we infer that efficient analyte breakdown largely occurs in the plasma rather than in the reaction tube even at relatively high aerosol carrier gas flow rates used in our experiments compared to typical values of ~1 SLPM in conventional ICP-MS.

84

Figure 3.2. Ratio of Cl response factors from flow injections of NaCl and chloramphenicol as a function of carrier gas flow rate at various exhaust velocity settings. Error bars reflect one standard deviation of the average based on triplicate injections of each compound.

85

3.3.2. Sampling Plasma Products into the Reaction Tube

Analyte breakdown products are carried by the argon gas from the central channel of the plasma into the reaction tube. Therefore, the fraction of the gas flow from the central channel that enters the reaction tube reflects the sampling efficiency of the plasma products. The central channel gas flow is in turn largely composed of the aerosol carrier gas flow, especially at high carrier gas flow rates [23, 24]. Accordingly, one can estimate the sampling efficiency using the ratio of the argon flow rate in the reaction tube to total aerosol carrier gas flow rate. Note that flow rates in standard conditions must be used for the ratios to account for drastic changes in temperature between the gas introduced into the plasma and the gas in reaction tube.

To determine the argon flow rate in the reaction tube, the gas velocity (v) and the temperature of the gas at the tube outlet were monitored as discussed in the Experimental section. Assuming a laminar flow with a parabolic profile for an ideal gas, the standard (273 K, and 1 atmosphere) volumetric gas flow rate (Fv) is calculated using Equation 3.1:

휋푟2푣 273 푃 퐹 = × × (eq 3.1) 푣 2 푇 1 where r is the inner radius of the reaction tube, T is the gas temperature at the tube end, and P is the lab pressure in atmospheres. The factor 2 in the denominator arises from the relationship between average gas velocity and the maximum velocity measured at the center of a laminar flow profile. The sampling efficiency is then calculated by Fv/Ft where Ft is the total carrier gas flow rate in standard liters per minute (SLPM) supplied by mass flow meters in the sample introduction.

Figures 3.3a and 3.3b show the effects of the total aerosol carrier gas and exhaust flow rates on the gas temperature and gas velocity at the reaction tube outlet. Figure 3.3a depicts that the reaction tube gas temperature generally increases with increasing carrier gas flow rate and

86 decreasing exhaust flow rate, illustrated by the diagonal trend from the upper left corner to the bottom right corner of Figure 3.3a. At lower exhaust flow rates, a larger flow of argon from the plasma is directed into the reaction tube. This provides a greater flow of thermal energy into the tube, causing higher temperatures in the reaction tube. The effect of aerosol carrier gas is determined by two opposing factors. As the aerosol gas flow rate increases, so does the amount of hot argon gas entering the reaction tube (see the sampling efficiency discussion below). This results in an increase in the reaction tube gas temperature. However, higher aerosol gas flow rate also leads to plasma cooling, counteracting the effect of greater fluid flow rate entering the tube.

The increasing reaction tube temperatures with aerosol gas flow rates up to 2.4 SLPM in Figure

3.3a suggest that the greater fluid flow into the tube is the dominant factor in this operating range.

Interestingly, the tube outlet gas temperature decreases at carrier gas flow rates > 2.4 SLPM for all exhaust settings. This indicates significant plasma cooling to the extent that the larger flow of argon into the tube does not compensate for the loss of thermal energy due to plasma cooling. This observation is also in agreement with deviations from compound-independent response factors at aerosol gas flow rates > 2.4 SLPM in Figure 3.2, further confirming the excessive plasma cooling at high aerosol flow rates that compromises analyte breakdown. Figure 3.3b illustrates that the reaction tube gas velocity also increases in a diagonal manner similar to tube gas temperature.

However, the impact of the exhaust flow rate on gas velocity is minimized as the carrier gas flow rate is increased, evident from near vertical contour lines at aerosol carrier gas flow rates > 2.4

SLPM in Figure 3.3b.

87

Figure 3.3. Effect of aerosol carrier gas flow rate and exhaust flow rate at 1100 W on a) temperature at reaction tube outlet, b) argon gas velocity flowing out of the reaction tube, c) sampling efficiency into the reaction tube calculated from the ratio of measured gas flow rate at the end of the reaction tube (Equation 3.1) to the total aerosol carrier gas flow rate, and d) Cl- sensitivity measured as the average peak area from triplicate injections of 12.5 μM chloramphenicol. Contour plots are utilized to facilitate identification of trends using Delaunay triangulation and linear interpolation between the data points to color the contours. Experimental values are indicated at the grid points. A nozzle-skimmer potential difference of 130 V was used in experiments of Figure 3.3d to improve Cl- detection sensitivity.

88

Figure 3.3c depicts the sampling efficiencies calculated from the data in Figures 3.3a and 3.3b using Equation 3.1. An increase in plasma sampling efficiency is observed with increasing aerosol gas flow rate and decreasing exhaust flow rate (diagonal trend) similar to Figures 3.3a and 3.3b.

The effect of exhaust velocity is minimized at aerosol gas flow rates > 2.4 SLPM, indicating the dominance of carrier gas flow rate in determining plasma sampling efficiency at high aerosol flow rates. Note that such high flow rates do not produce compound-independent response (Figure 3.2), thus are less desirable for quantitative elemental analysis. Interestingly, the sampling efficiencies are lower than 50% within the analytically preferred region for elemental quantification (aerosol gas flow rate < 2.4 SLPM) in Figure 3.3c, indicating that sensitivity improvements could be made by modification of the interface to sample a larger fraction of the central channel gas flow.

3.3.3. Effect of Chemical Reactions

With the insights gained above for the impact of operating parameters on analyte breakdown and plasma sampling efficiencies, we now focus on evaluating effects of chemical reactions on Cl- sensitivity. To this end, we examined the effect of aerosol carrier gas flow and exhaust flow rates on Cl- sensitivities in operating conditions similar to those in Figures 3.3a-3.3c. The results are shown in Figure 3.3d as a contour map for flow injection peak areas of a 12.5 μM chloramphenicol sample (25 μM Cl).

Figure 3.3d illustrates that the highest Cl- sensitivity is observed at total carrier gas flow rate of 2.0 SLPM and an exhaust velocity of 5.8 m/s. We expect near-complete vaporization and breakdown of analytes in these optimum operating conditions based on the compound-independent response shown in Figure 3.2. However, the trends for sensitivity in Figure 3.3d do not coincide well with trends for plasma sampling efficiency (Figure 3.3c) within the analytically preferred

89 region for quantitative elemental analysis (aerosol gas flow rate < 2.4 SLPM). In other words, plasma operating conditions that increase sampling efficiency do not necessarily lead to improved ion detection efficiencies, and the optimum ion signal is observed at a plasma sampling efficiency of only 26%. This observation highlights the major influence of complex chemical reactions on detected Cl- signals as discussed below.

Consider the operating conditions of 1.8 SLPM total carrier gas flow rate and exhaust velocity of 3.8 m/s (bottom left corner in Figure 3.3c). The sampling efficiency into the reaction tube in these conditions (28%) is similar to that of the optimum operating conditions for Cl- sensitivity

(26% at 2.0 SLPM carrier flow rate and 5.8 m/s exhaust velocity). Further, both reaction tube gas temperature and gas velocity in Figures 3.3a and 3.3b are similar between the bottom left corner and the optimum operating conditions. Therefore, species sampled into the reaction tube are likely to experience equivalent residence time and reaction temperature in the afterglow between the two conditions. Nevertheless, sensitivity in the bottom left corner of Figure 3.3d is 55 times lower than that at the optimum operating conditions. Such a drastic difference in sensitivity with similar plasma sampling and afterglow reaction conditions indicates a drastic change in the chemical nature of the species sampled into the reaction tube between the two conditions. The chemical transformations of the analytes in the plasma prior to sampling into the reaction tube is controlled by aerosol residence time within the plasma which in turn is a function of aerosol gas flow rate.

Therefore, variations in chemical composition of the plasma sampling point at different aerosol gas flow rates are expected. However, one must note that in-plasma reactions of the analytes at high aerosol gas flow rates in our experiments are likely more complex than mere atomization, and are likely to involve formation of polyatomic species as discussed later in this report.

90

Now consider a total carrier gas flow rate of 2.0 SLPM (optimum value for Cl sensitivity) at various exhaust velocities in Figure 3.3. In these conditions, the in-plasma reactions are set by the carrier gas flow rate and are expected to be independent of the exhaust flow rate. This expectation is further supported by the minimal impact of the exhaust flow rate on the response factor ratios in

Figure 3.2 at aerosol gas flow rate of 2.0 SLPM. Figure 3.3c shows that at constant carrier gas flow rate of 2.0 SLPM, the sampling efficiency increases slightly as the exhaust velocity is reduced from the optimum value of 5.8 m/s to 3.8 m/s. In contrast, the sensitivity is reduced by fivefold as depicted in Figure 3.3d. Notably, both velocity and temperature of the gas in the reaction tube increase as the exhaust velocity is reduced (Figures 3.3a and 3.3b). Therefore, the sampled species from the plasma spend less time in the reaction tube and the post-plasma reactions occur at a higher temperature. These results suggest that reactions in the tube also play a critical role in the ionization of Cl in addition to the in-plasma reactions controlled by the total carrier gas flow.

In summary, the results indicate that both in-plasma and post-plasma reactions play critical roles in Cl- detection. To gain insight into these reactions, we have explored the background ions and their relation to Cl-.

3.3.3.1. Background Ions

In our previous work on Cl- detection using ICP-based PARCI [19], four major background

- - - - ions were identified as NO2 , CN , OCN , and HCO2 upon introduction of methanolic sodium

- acetate into the plasma. Among these ions, HCO2 was specific to methanol introduction,

- - - suggesting that NO2 , CN , OCN are mainly a result of nitrogen diffusion into the plasma and afterglow while the formate ion is a product of methanol introduction. Importantly, we observed a

- - loose correlation between Cl sensitivity and HCO2 intensity, indicating a potential mechanistic

91 relation. These experiments, however, were conducted at high declustering potentials (nozzle- skimmer potential difference of 130 V) to improve Cl- ion detection. Such ion transmission conditions lead to loss of easily fragmentable species, hampering a comprehensive insight into the chemical composition of the ionization environment. In the current report, we collected spectra using low declustering potentials achieved at a nozzle-skimmer potential difference of 10 V to further the insights into ionization reactions.

Figure 3.4 compares the background spectra with high and low declustering potentials at optimum plasma operating conditions for Cl- detection. It is clear that high-mass ions observed by soft ion transmission (low declustering) are fragmented in harsh ion transmission conditions.

- - - Notably, [Na(HCO2)2] , [NaHCO2NO2] and [NaHCO2OCN] are the most prominent ions in the spectrum of Figure 3.4 with soft ion transmission. Molecular formulas of these clusters were further confirmed by exact mass measurements using a time-of-flight mass spectrometer with mass accuracy better than 5 ppm.

The presence and relative intensities of clusters in Figure 3.4 suggest the existence of NaHCO2

- - as a major species in the reaction tube to form first order clusters ([Na(HCO2)2] , [NaHCO2NO2]

- - - - and [NaHCO2OCN] ) via adduction to HCO2 , NO2 , and OCN as shown in thermodynamically

- - favorable Reactions 1-3 in Table 3.2. Higher cluster orders ([Nan(HCO2)n+1] , [Nan(HCO2)nNO2] ,

- and [Nan(HCO2)nOCN] , with n > 1) are formed by the stepwise reaction of sodium formate with the first order clusters. Note that the gas temperature in the tube is reduced during the transfer of species from the plasma toward the mass spectrometer. Calculations in Table 3.2 are conducted at a temperature of 600 K, representing the tube outlet temperature at optimum operating conditions.

92

Figure 3.4. Effect of declustering potential on background ions at optimized ion source operating conditions for best Cl- sensitivity (2.0 SLPM carrier flow rate, 5.8 m/s exhaust velocity, 1100 W RF power). The inset shows enlarged mass spectrum for the section with m/z >

170. Methanolic sodium acetate was used as ionization reagent. High declustering refers to 130 V potential difference between nozzle and skimmer of the MS while low declustering refers to 10 V potential difference.

93

Table 3.2. Reaction thermochemistries calculated at the ωB97xD/aug-cc-pVTZ level ΔH ΔG Number Reaction 600K 600K (kJ/mol) (kJ/mol) - - 1 HCO2 + NaHCO2  [Na(HCO2)2] -196 -127 - - 2 NO2 + NaHCO2  [NaHCO2NO2] -189 -120 - - 3 OCN + NaHCO2  [NaHCO2OCN] -156 -102 - - 4 Cl + NaHCO2  [NaHCO2Cl] -178 -128 - - 5 Cl + NaNO2  [NaNO2Cl] -191 -141 - - 6 NaCl + HCO2  [NaHCO2Cl] -226 -157 - - 7 NaCl + NO2  [NaNO2Cl] -219 -151 - - 8 NaCl + HCO2  Cl +NaHCO2 -48 -29 - - 9 NaCl + NO2  Cl + NaNO2 -27 -10 - - 10 NaCl + [Na(HCO2)2]  [NaHCO2Cl] + NaHCO2 -30 -30 - - 11 NaCl + [NaHCO2NO2]  [NaNO2Cl] + NaHCO2 -30 -30 - - 12 NaCl + [NaHCO2NO2] [NaHCO2Cl] + NaNO2 -17 -17 - - 13 F +NaHCO2  NaF + HCO2 -41 -57 - - 14 F + NaNO2 NaF + NO2 -62 -76 - - 15 NaF + HCO2  [NaHCO2F] -202 -134 - - 16 NaF + NO2  [NaNO2F] -196 -128 - - 17 NaF + [Na(HCO2)2]  [NaHCO2F] + NaHCO2 -6 -7 - - 18 NaF + [NaHCO2NO2]  [NaNO2F] + NaHCO2 -7 -8 - - 19 NaF + [NaHCO2NO2] [NaHCO2F] + NaNO2 +7 +6

3.3.3.2. Relation of Background Ions to Cl-

To identify the background ions relevant to Cl- detection, we measured the background spectra using soft ion transmission (low declustering) at a range of exhaust and carrier flow rates identical to those in Figure 3.3. We then correlated the results with the Cl- sensitivities of Figure 3.3d by linear regression of (x,y) points where x is the measured background ion intensity and y is the Cl- sensitivity at each operating condition. All background ions in the mass range 15-350 that showed intensities ≥ 100,000 cps in at least one operating condition were considered. Figure 3.5 shows the r2 of the regressions as a function of the ion m/z. No strong correlation was observed based on absolute ion intensities. However, when the ion intensities were normalized to the total ion count

(TIC) in each spectrum, strong correlations with Cl- sensitivity were observed for m/z 113 (r2 =

94

2 - 0.95) and 114 (r = 0.86) as shown in Figure 3.6. These m/z values correspond to [Na(HCO2)2]

- and [NaHCO2NO2] ions. Note that the carbon isotope contribution to m/z 114 from m/z 113 is

- minimal compared to the contribution from [NaHCO2NO2] as evident from the relative intensities of the two ions in Figure 3.4. Therefore, no isotopic correction is considered in this analysis. The correlations create a link between formate, nitrite, sodium and chloride and offer an avenue to explore the mechanistic role of sodium and methanol introduced into the plasma for enhanced Cl- detection.

Figure 3.7 depicts contour maps of TIC-normalized and absolute background ion intensities for nitrite, formate and their corresponding cluster ions. A clear visual similarity exists between

- - [Na(HCO2)2] TIC-normalized intensity map in Figure 3.7b and Cl sensitivity in Figure 3.3d (both having maximum values at aerosol gas flow = 2.0 SLPM and exhaust velocity of 5.8 m/s) while a

- lower degree of similarity is observed between [NaHCO2NO2] TIC-normalized intensity map

(Figure 3.7f) and Cl-. These visual correlations are consistent with the results of regression analyses in Figure 3.6. Further, the maps for absolute ion intensities in Figure 3.7 do not show similarities to the map of Cl- sensitivity consistent with regressions in Figure 3.5.

95

Figure 3.5. Correlation of background ion intensities with Cl- sensitivity indicated by r2 values for linear regressions. All ions with > 100000 cps intensity were considered in regressions.

96

Figure 3.6. a) r2 values for linear regression of background ions’ normalized intensities

(collected at low declustering nozzle-skimmer potential) with 35Cl- sensitivity (measured at high declustering potential) at carrier gas flow and exhaust flow rates identical to those in

- Figure 3.3. b) Correlation plots of normalized intensities for m/z 113 ([Na(HCO2)2] ), and c) m/z

- 114 ([NaHCO2NO2] ). The background ion intensities were normalized to the total ion counts in each spectrum.

97

Figure 3.7. a-h) TIC-normalized and, i-p) absolute intensities of formate and nitrite ions and their clusters with sodium formate as a function of carrier gas flow rate and exhaust setting.

The spectra were collected in soft ion sampling conditions using a low nozzle-skimmer declustering potential difference of 10 V.

98

The TIC-normalized values quantify the prominence of the ions within the spectra and reflect the importance of a particular ionization pathway relative to other reactions. For example,

- - [Na(HCO2)2] and [NaHCO2NO2] show their highest absolute intensities at an exhaust setting of

3.8 m/s and carrier gas flow rate of 2.4 SLPM in Figures 3.7j and 3.7n. However, the Cl- sensitivity in these conditions is 2.7 times lower than that at the optimum operating conditions (Figure 3.3d).

- - In contrast, TIC-normalized intensities for [Na(HCO2)2] and [NaHCO2NO2] are reduced by factors of 2 and 2.4 at exhaust setting of 3.8 m/s and carrier gas flow rate of 2.4 SLPM compared to the optimum operating conditions (Figures 3.7b and 3.7f). The reduced TIC-normalized intensities indicate increased contribution of other ionization reactions relative to the first-order cluster formation. Two examples of such reactions are revealed upon examination of the background spectrum at exhaust setting of 3.8 m/s and carrier gas flow rate of 2.4 SLPM shown

- in Figure 3.8: 1) new prominent ions compared to Figure 3.4 such as [(HCO2H)NO2] ,

- - - - [(HCO2H)HCO2] , [(HCO2H)OCN] , [Na(HCO2)2(HCO2H)] , and [Na(HCO2)NO2(HCO2H)] indicate presence of formic acid (HCO2H) and clustering of ions with this neutral as a major

- ionization pathway, and 2) enhanced intensities of second order clusters ([Na2(HCO2)3] and

- - - [Na2(HCO2)2NO2] ) relative to first-order clusters ([Na(HCO2)2] and [Na(HCO2)NO2] ) highlight a shift in cluster size distribution toward higher orders compared to the optimum conditions shown in Figure 3.4. This indicates promoted clustering with NaHCO2.

In summary, best Cl- sensitivities are observed in conditions that not only produce high

- - intensities for [Na(HCO2)2] and [Na(HCO2)NO2] but also ensure the favorable formation of theses first-order clusters over other reactions.

99

Figure 3.8. Background spectrum collected using soft ion sampling conditions (declustering potential of 10 V) at carrier gas flow rate of 2.4 SLPM and exhaust flow setting of 3.8 m/s.

Compared to Figure 3.4, new clusters are detected, indicating formation of HCO2H in these

- - conditions. Further, [Nan(HCO2)n+1] and [Nan(HCO2)nNO2] clusters with n > 1 are enhanced relative to clusters with n=1 compared to the data in Figure 3.4, showing a higher extent of clustering with HCO2Na.

100

3.3.3.3. Effect of Exhaust Flow and Carrier Gas Flow Rates on Clustering

- - The changes in cluster size distribution for [Nan(HCO2)n+1] and [Nan(HCO2)nNO2] offer insights into the impact of exhaust and carrier gas flow rates on clustering reactions. To examine these effects, we investigated the impact of operating parameters on average cluster order (A) for each family of cluster ions defined by Equations 3.2 and 3.3:

∑3 − 푛=0 푛×퐼 [푁푎푛(퐻퐶푂2)푛+1] 퐴 − = (eq 3.2) 퐻퐶푂2 3 − ∑푛=0 퐼 [푁푎푛(퐻퐶푂2)푛+1]

3 − ∑푛=0 푛×퐼[푁푎푛(퐻퐶푂2)푛푁푂2] 퐴 − = (eq 3.3) 푁푂2 3 − ∑푛=0 퐼[푁푎푛(퐻퐶푂2)푛푁푂2]

Figure 3.9 depicts the contour maps of the average cluster order. Note that the values at the aerosol gas flow rate of 1.8 SLPM in Figure 3.9 should be treated with caution because of the large uncertainties caused by low ion intensities at these operating parameters (see Figures 3.7i-p).

Figures 3.9a and 3.9b demonstrate that average cluster order increases with increasing exhaust setting at a constant aerosol gas flow rate of 2.0 SLPM (optimum for Cl- detection). This is in agreement with the trends in Figures 3.3a and 3.3b where reaction tube gas temperature and gas velocity are reduced with increasing exhaust flow at aerosol gas flow rate of 2.0 SLPM. Lower temperatures and longer residence times in the ICP afterglow promote clustering and shift the cluster order distribution to higher orders. Interestingly, the change in average cluster order as a function of the exhaust flow is minimized at higher aerosol gas flow rates. This trend is also in agreement with Figures 3.3a and 3.3b where the effect of exhaust flow on gas temperature and gas velocity are reduced at high aerosol gas flow rates.

101

Figure 3.9. Effect of aerosol carrier gas flow and exhaust flow rates on average cluster order

- - (n) for a) [Nan(HCO2)n+1] and b) [Nan(HCO2)nNO2] ions calculated using Equations 3.2 and

3.3.

102

Figures 3.9a and 3.9b also depict that the average cluster size generally increases as the aerosol gas flow rate is increased from the optimum value of 2.0 SLPM. This result is intriguing at first because one expects a lower clustering extent based on higher gas temperature and gas velocity within the reaction tube, evident from Figures 3.3a and 3.3b for increasing aerosol gas flow rates.

However, it must be noted that plasma conditions, upstream of the reaction tube are also impacted by higher carrier gas flow rates. Specifically, increased carrier gas flow rates reduce the residence time of aerosol in the plasma and decrease the temperature of the central channel. In-plasma reactions in such operating conditions are expected to promote the formation of polyatomic species, like NaHCO2. The ensuing increase in the concentration of NaHCO2 within the reaction tube could ultimately shift the clustering reactions (Reactions 1-3, Table 3.2) to the right, compensating for the opposite effects of higher temperature and shorter afterglow reaction times.

3.3.3.4. Implications for Cl Ionization

- - To elucidate the relationship between the prominence of [Na(HCO2)2] and [Na(HCO2)NO2] clusters in the spectra and chloride ion detection, we examined the possibility of detecting similar

Cl-containing clusters. Accordingly, we monitored ions at m/z values corresponding to Cl-,

- - - - [NaHCO2Cl] , [NaNO2Cl] , [Na2(HCO2)2Cl] , and [Na2HCO2NO2Cl] using SIM mode upon flow injections of chloramphenicol into the plasma with aerosol gas flow rate of 2.0 SLPM, exhaust velocity of 5.8 m/s and RF power of 1100 W. These conditions correspond to the highest signal for Cl- (Figure 3.3d). However, unlike the experiments for Figure 3.3d, we utilized a low declustering potential (10 V) to preserve the clusters.

103

Figure 3.10. Relative sensitivities (flow injection peak areas) of chlorine-containing ions detected in soft ion sampling conditions (nozzle-skimmer potential difference of 10 V) upon injection of chloramphenicol. Sensitivities are normalized to that of the most intense chlorinated

35 - ion ([NaHCO2 Cl] ). Error bars indicate one standard deviation of the average based on triplicate injections.

104

Figure 3.10 shows relative intensities for the expected species, confirming their formation.

Notably, the distribution of Cl-containing species in Figure 3.10 resembles those of

- - [Nan(HCO2)n+1] and [Nan(HCO2)nNO2] clusters (See Figure 3.4) where the first-order clusters are the dominant species. The larger abundance of the Cl-containing clusters in Figure 3.10 relative to the un-clustered Cl- suggests that the Cl- ions detected in harsh ion transmission (e.g. in Figure

3.3d) are largely a result of cluster ion breakdown upon ion activation in the nozzle-skimmer area.

The calculated thermochemistries in Table 3.2 can be used to evaluate the competitive

- fragmentation pathways upon activation of the cluster ions. For example [Na2(HCO2)2Cl] and

- - - - [Na2HCO2NO2Cl] may release Cl , HCO2 , and NO2 upon activation. Thermochemistries for

Reactions 4-7 in Table 3.2 show that the release of Cl- from the first-order clusters requires less

- - - energy compared to that needed for release of NO2 and HCO2 . Therefore, Cl is predicted to be the major fragmentation product of the clusters.

Experimental investigation of the fragmentation products require MS/MS experiments.

Coupling PARCI to an MS/MS instrument is currently under development in our laboratory.

Fortunately, similar clusters can be generated using electrospray ionization from a solution of sodium chloride and sodium formate as shown in Figure 3.11a, obtained using an ESI-triple

- quadrupole MS instrument. Figure 3.11b depicts that activation of [Na(HCO2)Cl] at a collision

- - energy (CE) of 12 eV mainly produces Cl as predicted above. Release of HCO2 becomes competitive at higher activation energies (CE = 20 eV, Figure 3.11c) where fragmentation channels with higher activation barriers are accessed.

105

106

Figure 3.11. a) MS spectrum of clusters generated by nano-ESI from an aqueous solution of

- sodium chloride and sodium formate, b) MS/MS spectrum of [NaHCO2Cl] at collision

- - energy of 12 eV showing Cl as a major fragment ion, c) MS/MS spectrum of [NaHCO2Cl] at collision energy of 20 eV, indicating competitive generation of formate and chloride at

- higher collision energies, d) MS/MS spectrum of [Na2(HCO2)2Cl] at collision energy of 20 eV, showing loss of NaHCO2 as the major fragmentation pathway from the second order

- cluster, e) MS/MS spectrum of [Na3(HCO2)3Cl] at collision energy of 20 eV, showing successive loss of NaHCO2 as the major fragmentation pathway from the third order cluster.

Experiments were performed using a Sciex API3000 instrument with a homebuilt nanospray source composed of a pulled glass capillary with a tip size of 3 μm. An aqueous solution of 1 mM

NaHCO2 and 100 μM NaCl was loaded into the capillary. A platinum wire was inserted into the capillary to apply -1100 V to the solution while a voltage of -200 V was applied to the curtain plate of the instrument using an external power supply. Nitrogen was used as CAD gas at an instrument setting of 4.

107

Importantly, the cluster order (n) has a major influence on the efficiency of Cl- release from

- - [Nan(HCO2)nCl] and [Nan(HCO2)n-1NO2Cl] clusters. As the cluster order grows, more fragmentation pathways become available, competing for the energy deposited in the ion during

- activation. This is shown in Figures 3.11d and 3.11e for fragmentation of [Na2(HCO2)2Cl] and

- [Na3(HCO2)3Cl] at CE = 20 eV, demonstrating that lower order clusters become the major

- fragmentation products via loss of NaHCO2. The reduced fragmentation efficiency to Cl for cluster orders (n) > 1 is in agreement with the lower Cl- sensitivity observed at operating conditions that promote clustering beyond the first-order clusters.

3.3.3.5. Mechanistic Considerations for Enhancement of Cl- Detection by First-Order

Clusters

To understand the role of methanol (responsible for formate generation) and sodium in improving Cl- detection from ICP afterglow, we now consider two ionization mechanisms that involve formation of first order clusters. In the first mechanism, Cl- ions form in the plasma or in the afterglow via electron capture reactions and react with plasma generated neutrals in the reaction tube to create the chlorinated clusters. Reactions 4 and 5 in Table 3.2 indicate favorable thermochemistry for clustering reactions of Cl-. In this mechanism, improvements in Cl- detection upon introduction of sodium and methanol could be explained by preventing Cl- loss upon clustering. One possibility for a Cl- loss mechanism is ion-ion recombination in the afterglow.

Clusters will have lower mobilities compared to the Cl- ion, reducing the ion-ion recombination rate.

The second mechanism is formation of NaCl and subsequent ion-neutral reactions. One can consider sodium transfer reactions of NaCl with formate and nitrite, leading to release of Cl- as

108 shown by Reactions 8 and 9 in Table 3.2. Subsequently, the released Cl- could undergo clustering reactions as discussed above. Alternatively, clusters can form via adduction of formate and nitrite to NaCl (Reactions 6 and 7), or via formate and nitrite transfer from first order clusters of these ions (Reactions 10-12) in Table 3.2. Given the prominence of the first order clusters for formate and nitrite in our optimum experimental conditions for Cl- sensitivity (Figure 3.4), the formate and nitrite ions are likely to react rapidly with sodium formate, depleting the population of these ions in the reactive flow. Accordingly, we infer that the nitrite and formate transfer reactions (Reactions

- - 10-12 in Table 3.2) would play significant roles in formation of [NaHCO2Cl] and [NaNO2Cl] among the Reactions 6-12 which consider NaCl as an intermediate species in ionization of chlorine.

The major difference between the first and second mechanism is that the first one starts with a

Cl- ion and solely relies on loss prevention to explain sensitivity improvements by clustering, while the second one considers formation of NaCl and subsequent ionization of this neutral as an ionization route to improve formation of Cl- ions. Elucidation of relative importance between the two mechanisms would require further investigations, however, both mechanisms clearly indicate the importance of cluster formation in improving Cl- detection in the ICP afterglow.

3.3.3.6. Application to Fluorine Detection

In an effort to evaluate the effectiveness of above ionization reactions for analysis of fluorine, we investigated detection of fluoride ion and its clusters. Early studies in negative mode ICP-MS have revealed that Cl- and F- ions are observed with similar sensitivities as a result of complete atomization in the plasma and subsequent electron capture by the atoms [25–28]. In our experimental investigations, however, we have not been able to observe fluoride or fluorine-

109 containing clusters upon injection of 500 μM para-fluoro-phenylalanine while chlorine clusters are readily observed by injections of 10 μM chloramphenicol and 20 μM NaCl. This observation is consistent with both mechanisms discussed above. F- has a large sodium ion affinity and can readily undergo sodium ion transfer reaction with sodium formate and sodium nitrite forming NaF

(Reactions 13 and 14). Accordingly, both mechanisms predict NaF as a plasma product for fluorinated compounds. The adduction of nitrite and formate to NaF is favorable (Reactions 15 and 16) and can be an ionization route for cluster formation. However, formate and nitrite ions are rapidly depleted via adduction to NaHCO2 in our experimental conditions as discussed above, making the formate and nitrite transfer reactions (17, 18, 19) dominant ion formation pathways.

Importantly, the thermochemistries indicate that the anion transfer to NaF is not as favorable as that for NaCl, likely resulting in establishment of equilibrium for these reactions. Moreover, high concentration of NaHCO2 would push the equilibrium left in reactions 17-19, further reducing ionization efficiency. In summary, the ionization mechanisms discussed above are consistent with the experimental observations for both F and Cl, offering support for the explanation of chloride ion enhancement in the presence of methanol and sodium, while fluorine detection as a negative ion is hindered in similar experimental conditions.

3.4. Conclusions

We have examined the major chloride ion formation steps for elemental analysis of chlorine using an Ar-ICP with an atmospheric afterglow reaction region. This configuration is an implementation of a general approach described as plasma-assisted reaction chemical ionization.

Our results indicate that aerosol gas flow rates ≤ 2.2 SLPM using a 1.5-mm i.d. injector and 1100

W RF power provide sufficient breakdown of analytes for compound-independent Cl response.

110

Unlike conventional ICP-MS, the exhaust flow rate plays a critical role in high-sensitivity detection of Cl in plasma-assisted reaction chemical ionization. The effect of exhaust is two-fold:

1) it controls the sampling efficiency of plasma products into the reaction tube, and 2) it affects the temperature and velocity of the gas in the reaction tube, in turn impacting ionization reactions.

The gas flow measurements in the tube reveal that sampling efficiencies of ~26% are achieved at optimum operating conditions for best Cl- sensitivity, indicating that sensitivity enhancements could be achieved by improved reaction interface designs with higher plasma sampling efficiencies.

Importantly, key observations have revealed cluster formation as a critical process for detection of Cl- using methanolic sodium acetate as an ionization reagent in PARCI. These observations include: 1) strong correlation between Cl- sensitivities (measured at high declustering potentials)

- - and TIC-normalized intensities of [Na(HCO2)2] and [NaHCO2NO2] detected with soft ion

- - sampling, and 2) observation of [NaHCO2Cl] and [NaNO2Cl] as dominant chlorine-containing ions detected with soft ion sampling (low declustering potential) upon injection of chlorinated analytes. These observations indicate that the improved Cl- detection at high declustering potentials is likely a result of cluster fragmentation and release of Cl- from various clusters upon ion activation in the nozzle-skimmer area of the MS. The formation of clusters is influenced by the aerosol carrier gas flow rate, which determines the in-plasma reaction conditions as well as the gas temperature and velocity in the reaction tube. Similarly, the torch box exhaust setting affects the cluster formation via its influence on reaction tube gas temperature and gas velocity (reaction time).

111

The chlorine-containing clusters may be a result of chloride ion formation and subsequent adduction to plasma generated neutrals (such as NaHCO2), preventing chloride ion loss via ion- ion recombination reactions. Alternatively, the clusters may be generated via formation of NaCl and transfer of formate and nitrite from first order clusters of these ions. Both mechanisms are viable as indicated by ab-initio calculations, and offer an explanation for improved Cl- detection via concurrent introduction of sodium and methanol into the plasma. These two mechanisms are summarized in the schematic diagram of Figure 3.12. Note that formate is a product of methanol introduction. Both mechanisms are also consistent with the low sensitivity of fluorine detection in negative mode PARCI-MS.

Our studies indicate that effective atmospheric-pressure ion-molecule reactions can be utilized to increase sensitivity of elemental ionization for chlorine. In particular, the insights gained in these studies provide the grounds for designing a more sensitive ion source for detection of Cl and F. Importantly, implementation of the afterglow reactions also creates a low- temperature buffer between the extremely hot ICP and the mass spectrometer. Accordingly, a wide range of commercially available atmospheric-sampling mass spectrometers with high- resolution and MS/MS capabilities could be used for PARCI-MS to improve analytical performance.

112

Plasma Reaction Droplet Tube MS

- [Na(HCO2)2] NaCl CH3OH - + CH CO 3 2 + 113 Na - R-Cl [NaHCO2Cl] - or Cl + - NaCHO2 Cl NaHCO2

+

Figure 3.12. Schematic diagram summarizing the proposed Cl- ionization mechanisms. Droplets containing organochlorine

analytes, sodium, and methanol are introduced to the plasma where analytes are converted to Cl- directly or to NaCl leading to reactions

- - - with [Na(HCO2)2] or NaHCO2 derived from sodium and methanol. These reactions form [NaHCO2Cl] , which yields Cl upon

activation.

113

3.5. References

1. Niu, H., Houk, R.S.: Fundamental aspects of ion extraction in inductively coupled plasma mass spectrometry. Spectrochim. Acta Part B At. Spectrosc. 51, 779–815 (1996). https://doi.org/10.1016/0584-8547(96)01506-6

2. Secretariat of the Stockholm Convention: The Stockholm Convention on Persistent

Organic Pollutants, http://chm.pops.int/

3. Wishart, D.S., Feunang, Y.D., Guo, A.C., Lo, E.J., Marcu, A., Grant, J.R., Sajed, T.,

Johnson, D., Li, C., Sayeeda, Z., Assempour, N., Iynkkaran, I., Liu, Y., Maciejewski, A., Gale, N.,

Wilson, A., Chin, L., Cummings, R., Le, D., Pon, A., Knox, C., Wilson, M.: DrugBank 5.0: a major update to the DrugBank database for 2018. Nucleic Acids Res. 46, D1074–D1082 (2018). https://doi.org/10.1093/nar/gkx1037

4. US Food & Drug Administration: Pesticide Residue Monitoring Program, https://www.fda.gov/Food/FoodborneIllnessContaminants/Pesticides/ucm582707.htm

5. Lavonen, E.E., Gonsior, M., Tranvik, L.J., Schmitt-Kopplin, P., Köhler, S.J.: Selective

Chlorination of Natural Organic Matter: Identification of Previously Unknown Disinfection

Byproducts. Environ. Sci. Technol. 47, 2264–2271 (2013). https://doi.org/10.1021/es304669p

6. Remucal, C.K., Manley, D.: Emerging investigators series: the efficacy of chlorine photolysis as an advanced oxidation process for drinking water treatment. Environ. Sci. Water Res.

Technol. 2, 565–579 (2016). https://doi.org/10.1039/C6EW00029K

7. Kinani, A., Salhi, H., Bouchonnet, S., Kinani, S.: Determination of adsorbable organic halogens in surface water samples by combustion–microcoulometry versus combustion–ion

114 chromatography titration. J. Chromatogr. A. 1539, 41–52 (2018). https://doi.org/10.1016/j.chroma.2018.01.045

8. Klencsár, B., Bolea-Fernandez, E., Flórez, M.R., Balcaen, L., Cuyckens, F., Lynen, F.,

Vanhaecke, F.: Determination of the total drug-related chlorine and bromine contents in human blood plasma using high performance liquid chromatography–tandem ICP-mass spectrometry

(HPLC–ICP-MS/MS). J. Pharm. Biomed. Anal. 124, 112–119 (2016). https://doi.org/10.1016/j.jpba.2016.02.019

9. Klencsár, B., Balcaen, L., Cuyckens, F., Lynen, F., Vanhaecke, F.: Development and validation of a novel quantification approach for gradient elution reversed phase high-performance liquid chromatography coupled to tandem ICP-mass spectrometry (RP-HPLC-ICP-MS/MS) and its application to diclofenac and its related compounds. Anal. Chim. Acta. 974, 43–53 (2017). https://doi.org/10.1016/j.aca.2017.04.030

10. Barbosa, J.T.P., Santos, C.M.M., Bispo, L. dos S., Lyra, F.H., David, J.M., Korn, M. das

G.A., Flores, E.M.M.: Bromine, Chlorine, and Iodine Determination in Soybean and its Products by ICP-MS After Digestion Using Microwave-Induced Combustion. Food Anal. Methods. 6,

1065–1070 (2013). https://doi.org/10.1007/s12161-012-9511-6

11. Nelson, J., Hopfer, H., Silva, F., Wilbur, S., Chen, J., Shiota Ozawa, K., Wylie, P.L.:

Evaluation of GC-ICP-MS/MS as a New Strategy for Specific Heteroatom Detection of

Phosphorus, Sulfur, and Chlorine Determination in Foods. J. Agric. Food Chem. 63, 4478–4483

(2015). https://doi.org/10.1021/jf506372e

115

12. Bu, X., Wang, T., Hall, G.: Determination of halogens in organic compounds by high resolution inductively coupled plasma mass spectrometry (HR-ICP-MS). J. Anal. At. Spectrom.

18, 1443–1451 (2003). https://doi.org/10.1039/B306570G

13. Gois, J.S. de, Costas-Rodríguez, M., Vallelonga, P., G. Borges, D.L., Vanhaecke, F.: A simple method for high-precision isotopic analysis of chlorine via pneumatic nebulization multi- collector inductively coupled plasma-mass spectrometry. J. Anal. At. Spectrom. 31, 537–542

(2016). https://doi.org/10.1039/C5JA00408J

14. Renpenning, J., Horst, A., Schmidt, M., Gehre, M.: Online isotope analysis of 37Cl/35Cl universally applied for semi-volatile organic compounds using GC-MC-ICPMS. J. Anal. At.

Spectrom. 33, 314–321 (2018). https://doi.org/10.1039/C7JA00404D

15. Ohata, M., Wada, A.: The determination of chlorine in plastic using isotope dilution inductively coupled plasma mass spectrometry. Anal. Methods. 9, 4004–4010 (2017). https://doi.org/10.1039/C7AY01065F

16. Wang, H., Lin, N., Kahen, K., Badiei, H., Jorabchi, K.: Plasma-Assisted Reaction

Chemical Ionization for Elemental Mass Spectrometry of Organohalogens. J. Am. Soc. Mass

Spectrom. 25, 692–695 (2014). https://doi.org/10.1007/s13361-013-0816-5

17. Wang, H., Minardi, C.S., Badiei, H., Kahen, K., Jorabchi, K.: High-sensitivity elemental ionization for quantitative detection of halogenated compounds. Analyst. 140, 8177–8185 (2015). https://doi.org/10.1039/C5AN01958C

18. Lin, N., Wang, H., Kahen, K., Badiei, H., Jorabchi, K.: Gas Chromatography Plasma-

Assisted Reaction Chemical Ionization Mass Spectrometry for Quantitative Detection of Bromine in Organic Compounds. Anal. Chem. 86, 7954–7961 (2014). https://doi.org/10.1021/ac501964u

116

19. Lesniewski, J.E., McMahon, W.P., Zheng, K., Wang, H., Badiei, H., Jorabchi, K.:

Atmospheric pressure plasma assisted reaction chemical ionization for analysis of chlorinated compounds separated by liquid chromatography. J. Anal. At. Spectrom. 32, 1757–1765 (2017). https://doi.org/10.1039/C7JA00115K

20. Hyttinen, N., Otkjær, R.V., Iyer, S., Kjaergaard, H.G., Rissanen, M.P., Wennberg, P.O.,

Kurtén, T.: Computational Comparison of Different Reagent Ions in the Chemical Ionization of

Oxidized Multifunctional Compounds. J. Phys. Chem. A. 122, 269–279 (2018). https://doi.org/10.1021/acs.jpca.7b10015

21. Hanwell, M.D., Curtis, D.E., Lonie, D.C., Vandermeersch, T., Zurek, E., Hutchison, G.R.:

Avogadro: an advanced semantic chemical editor, visualization, and analysis platform. J.

Cheminformatics. 4, 17 (2012). https://doi.org/10.1186/1758-2946-4-17

22. Frisch, M.J., Trucks, G.W., Schlegel, H.B., Scuseria, G.E., Robb, M.A., Cheeseman, J.R.,

Scalmani, G., Barone, V., Petersson, G.A., Nakatsuji, H., Li, X., Caricato, M., Marenich, A.V.,

Bloino, J., Janesko, B.G., Gomperts, R., Mennucci, B., Hratchian, H.P., Ortiz, J.V., Izmaylov,

A.F., Sonnenberg, J.L., Williams, Ding, F., Lipparini, F., Egidi, F., Goings, J., Peng, B., Petrone,

A., Henderson, T., Ranasinghe, D., Zakrzewski, V.G., Gao, J., Rega, N., Zheng, G., Liang, W.,

Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y.,

Kitao, O., Nakai, H., Vreven, T., Throssell, K., Montgomery Jr., J.A., Peralta, J.E., Ogliaro, F.,

Bearpark, M.J., Heyd, J.J., Brothers, E.N., Kudin, K.N., Staroverov, V.N., Keith, T.A., Kobayashi,

R., Normand, J., Raghavachari, K., Rendell, A.P., Burant, J.C., Iyengar, S.S., Tomasi, J., Cossi,

M., Millam, J.M., Klene, M., Adamo, C., Cammi, R., Ochterski, J.W., Martin, R.L., Morokuma,

K., Farkas, O., Foresman, J.B., Fox, D.J.: Gaussian 16 Rev. B.01. , Wallingford, CT (2016)

117

23. Lindner, H., Murtazin, A., Groh, S., Niemax, K., Bogaerts, A.: Simulation and

Experimental Studies on Plasma Temperature, Flow Velocity, and Injector Diameter Effects for an Inductively Coupled Plasma. Anal. Chem. 83, 9260–9266 (2011). https://doi.org/10.1021/ac201699q

24. Lindner, H., Bogaerts, A.: Multi-element model for the simulation of inductively coupled plasmas: Effects of helium addition to the central gas stream. Spectrochim. Acta Part B At.

Spectrosc. 66, 421–431 (2011). https://doi.org/10.1016/j.sab.2011.04.007

25. Fulford, J.E., Quan, E.S.K.: Negative Ions in Inductively Coupled Plasma Mass

Spectrometry. Appl. Spectrosc. 42, 425–428 (1988)

26. Vickers, G.H., Wilson, D.A., Hieftje, G.M.: Detection of negative ions by inductively coupled plasma mass spectrometry. Anal. Chem. 60, 1808–1812 (1988). https://doi.org/10.1021/ac00168a031

27. Chtaib, M., Schmit, J.-P.: Negative-ion mode in inductively coupled plasma mass spectrometry. J. Anal. At. Spectrom. 3, 315–318 (1988). https://doi.org/10.1039/JA9880300315

28. Pupyshev, A.A., Surikov, V.T.: Application of negative ions in inductively coupled plasma-mass spectrometry. Spectrochim. Acta Part B At. Spectrosc. 59, 1021–1031 (2004). https://doi.org/10.1016/j.sab.2004.05.011

118

Chapter IV

High-Sensitivity Elemental Mass Spectrometry of Fluorine by Ionization in Plasma

Afterglow‡

4.1. Introduction

Fluorinated compounds have become widely used chemicals. For instance, about 25% of agrochemicals, approximately 20% of pharmaceuticals, and 30% of the top selling drugs contain

F atoms [1, 2]. Further, oil- and water-repellent properties of per- and poly-fluorinated compounds have made them attractive components in fire retardant formulations [3], textile coatings [4], and food packaging liners [5], resulting in environmental and food contamination. Importantly, chemical and metabolic transformations of fluorochemicals play critical roles in determining the environmental fate of these compounds [6, 7], and their use as drugs [8]. As such, the need for analytical techniques to detect and quantify fluorochemicals and their transformation products has risen with time [9].

Notably, natural fluorinated compounds are rare, despite the abundance of fluorine on earth

[10]. Therefore, the fluorine atom offers an intrinsic tag to detect fluorochemicals in complex matrices. Elemental mass spectrometry (MS) is an attractive approach for elemental tag detection because it provides compound-independent response factors and high matrix tolerance, enabling quantitation of compounds without compound-specific standards. This capability is particularly useful when standards are not readily available, e.g. in non-targeted approaches where new

‡ This chapter has been modified, integrated with supplementary information, and reprinted with permission (see Appendix) from Lesniewski, J.E., Zheng, K., Lecchi, P., Dain, D., and Jorabchi, K.: High-Sensitivity Elemental Mass Spectrometry of Fluorine by Ionization in Plasma Afterglow. Anal. Chem. 91, 3773-3777 (2019). Copyright 2019 American Society of Chemistry. The authors would like to acknowledge that this material is based upon work supported by the National Science Foundation (NSF) under CHE-1507304. The authors thank PerkinElmer Inc. for the loan of the mass spectrometer and the ICP generator used in these studies. The authors thank Elemental Scientific Inc. for the gift of the aerosol mixer. The authors are grateful to Thermo Fisher Scientific for the loan of the ion chromatography unit. The authors would also like to note that Kaveh Jorabchi is an inventor on a US patent for PARCI-MS. 119 compounds are discovered. Further, elemental detection facilitates discovery of species in complex samples by flagging the retention times of analytes containing an element of interest within a chromatogram [11]. This attribute is useful in analysis of compounds with mono-isotopic heteroatoms (e.g. F) whose detection becomes difficult using molecular MS techniques in the absence of defining isotopic patterns.

Recent development of inductively coupled plasma (ICP)-MS/MS has improved analysis of compounds containing Cl, Br, S, and P by reducing isobaric interferences [12, 13]. However, elemental detection of fluorinated compounds has remained a major challenge because of fundamental ionization limitations in ICP-MS. Formation of F+ occurs with extremely poor efficiency in an Ar-ICP, resulting in low sensitivities [14]. Alternative approaches such as F- detection [15] and formation of BaF+ in the ICP [16–19] have been investigated to resolve this challenge, but have not yielded sensitive and robust ionization methods. Continuum source molecular absorption spectroscopy has also been investigated for element-targeted quantitation of fluorinated compounds by formation of GaF in a furnace atomizer, however, the approach is not compatible with online chromatographic detection [20, 21].

Here, we report a new ionization method for elemental MS of fluorine, offering compound- independent response factors and two orders of magnitude improved sensitivity compared to those reported for ICP-MS. Our approach is based on plasma assisted reaction chemical ionization

(PARCI) developed recently using an ICP as the plasma source [22, 23]. Further, we demonstrate quantitation of fluoride in infant formula using ion chromatography coupled to the new ionization technique as an example for elemental fluorine analysis of samples with complex matrices.

4.2. Experimental Details

120

a exhaust flow glass ion transfer capillary

atmospheric-pressure steel reaction tube

skimmer

NaF + + Na2F Na quadrupole

argon plasma

vacuum pumps cooled aluminum sampler ion attraction lens N2 counterflow gas 5.5 L/min

b exhaust flow sampler skimmer collision cell

18 + 18 + + Ba OH Ba OH + BaF BaF + BaF+ + O BaOH BaF 2

argon plasma vacuum pumps

supersonic expansion

Figure 4.1. Schematics of a) ICP-based PARCI-MS, and b) conventional ICP-MS/MS for detection of F.

121

Figure 4.1a depicts the major components of the instrumental setup where an Ar ICP operated at 1100 W is coupled to a 5-cm atmospheric-pressure reaction tube placed 5-10 mm from the inlet of an atmospheric-sampling single-quadrupole mass spectrometer.

4.2.1. Sample Introduction and PARCI

Aqueous analytes were introduced by flow injection using a 20 μL injection loop at 250 μL/min water supplied by an HPLC pump (LC10AD, Shimadzu, Columbia, MD). The analyte stream was then mixed with 80 μL/min 10 mM aqueous NaOH solution supplied by a syringe pump (Model

100, KD Scientific, Holliston, MA). The mixed stream was then guided into a nebulizer (HEN-

120, Meinhard, Golden, CO) operated at a constant argon flow rate of 1.25 L/min. The aerosols passed through a cyclonic spray chamber (Twister Cyclonic, Glass Expansion, Pocasset, MA) and were mixed with a makeup gas flow of 0.55-1.25 L/min using a tangential mixer (ML151008N,

Meinhard, Golden, CO) to provide total aerosol gas flow rates of 1.8-2.5 L/min.

The total aerosol gas flow was injected via a 1.5-mm injector into an Ar ICP sustained at 1100

W (14 L/min outer gas flow, 1.2 L/min auxiliary flow) using a stand-alone RF generator (Nexion

2000, PerkinElmer Inc., Waltham, MA) with the exhaust flow adjusted to produce 3.5 m/s air flow at the bottom intake of the torch box. Note that exhaust flow has a significant impact in PARCI unlike conventional ICP-MS [23]. Analytes underwent vaporization and breakdown within the plasma. The plasma products were then guided through a cooled aluminum sampler (2.5 mm orifice) into a stainless steel atmospheric-pressure reaction tube (5 cm long, ¼” o.d., 4.8 mm i.d.) where ion-molecular reactions led to ionization of plasma products.

4.2.2. Mass Spectrometer

The mass spectrometer used for detection of ions was a single quadrupole MS (SQ 300,

PerkinElmer Inc., Waltham, MA). The electrospray ion source was removed from the instrument

122 and the PARCI source was placed so that the reaction tube was aligned with the inlet, 5-10 mm from the MS end plate. Neutrals were restricted from entering the mass spectrometer by a 5.5

L/min N2 counterflow gas controlled by an external mass flow controller (Model 246B, MKS,

Andover, MA). Positive ions were pulled in to the MS inlet by applying potentials of -200 V to the end plate and -600 V to the glass capillary entrance. Background scans were collected using

1000 μs pulse counting time per point, 10 points per mass, and were averaged for 30 seconds. A nozzle-skimmer declustering potential of 50 V was used for all acquisitions (+70 V capillary exit,

+ +20 V skimmer). Na2F sensitivity measurements were conducted in selected ion monitoring

(SIM) mode at an m/z value of 65 with an integration time of 500 ms. For measurements of LOD, an integration time of 1000 ms was utilized. To minimize premature aging of the detector due to the high background ion flux, an intentional ion deflection was applied to reduce the ion current impinging on the detector. This was accomplished by setting the Q1 offset parameter to zero

(normally it is set to -2 V). The signal suppression factor was measured using flow injections with sequential suppressed and unsuppressed ion flux settings. The ion intensities measured at suppressed flux conditions were corrected using the measured factor for all of the values reported in this work.

4.2.3. Sample Preparation and Data Analysis

All solvents were LC-MS grade (Chromasolv, Sigma-Aldrich, St. Louis, MO). p-fluoro- phenylalanine (p-F-phe) was purchased from Sigma-Aldrich, St. Louis, MO and sodium fluoride was purchased from the JT Baker Chemical Company, Phillipsburg, NJ. Analyte solutions were prepared by dissolving solids in 2-mM NaOH in water at concentrations of ~100 mM for NaF and

~15 mM for p-F-phe followed by dilution in water to final concentrations of 6-500 μM fluorine in water. ACS grade sodium hydroxide (VWR, Radnor, PA) and sodium acetate (Fisher, Waltham,

123

MA) were used for post-column sodium addition. Sodium salt solutions were prepared by dissolving the solids in water at a concentration of ~2 M, followed by dilution to 10 mM in water.

All selected ion chromatograms from flow injections were exported as text files, and then manually integrated using the trapezoid rule by selecting the beginning and end points of a baseline for each peak. Peak heights were calculated as the difference between the top of the peak and the baseline.

4.2.4. Ab Initio Calculations

Gaussian 16 [24] was used to calculate reaction thermochemistries at a theory level of

ωB97xD/aug-cc-pVTZ at 600 K (the approximate temperature at the end of the reaction tube).

This level of theory has previously been used to describe the thermochemical properties of halide- neutral adducts [25]. Structures were optimized using the MMFF94 force field in the Avogadro program [26] to generate starting geometries for each species prior to optimization at the final theory level. For molecules where more than one conformer is possible, the lowest energy conformer was used.

4.2.5. Ion Chromatography

For measurements of F- in infant formula, an ion chromatography unit (ICS-2000 pump, 300- uL injection loop, an IonPac AG11-HC 4x50 mm guard column, and IonPac AS-11HC 4x250 mm anion exchange column, Thermo Fisher Scientific, Waltham, MA) with NaOH gradient elution was utilized. Separations occurred at a flow rate of 1.1 mL/min and the eluent was split after the conductivity detector with 250 uL/min mixed with 80 uL/min 10 mM NaOH prior to introduction into PARCI-MS. This arrangement allowed simultaneous conductivity and PARCI-MS detection.

Use of NaOH rather than other sodium salts was preferred to increase the pH of the solution after the column and to prevent reactions and loss of F- in transfer to the plasma.

124

PARCI-MS was operated at a total aerosol carrier gas flow rate of 2.2 L/min (1.25 L/min nebulizer, 0.95 L/min auxiliary) while other parameters were kept constant as described above.

Fluoride calibration curve was constructed using injections of NaF standard in 18 MΩ water from

6.7 to 122 µM of fluoride using a 5 mM NaOH isocratic elution.

For infant formula analysis, 1.5 g of powder sample (Enfamil, Mead Johnson & Company,

LLC, Evansville, IN) was dissolved in 12 ml of 18 MΩ water. The sample was subjected to centrifugal filtration using 3000 Da filters (Amicon Ultra-15, Millipore Sigma, Burlington, MA) to separate the proteins and large fats. Spiked infant formula samples were prepared by adding 15 and 52 µM fluoride to 10 ml of infant formula sample using a NaF standard solution prior to filtration. The filtrate was directly injected into ion chromatography without any further sample treatment. Separation of fluoride in infant formula was achieved via a multi-step NaOH gradient starting at 0.2 mM for 4 min followed by 5 mM for 16 min. The eluent from the IC was connected to PARCI only between elution times of 5 and 10 min for infant formula samples to avoid introduction of unretained and strongly retained matrix into PARCI. The column was cleaned using 80 mM NaOH for 10 min followed by reconditioning at 0.2 mM NaOH for 10 min prior to each injection.

+ 4.3. Na2F as a New Reporter Ion for F

Analytes are introduced into the ICP where they form element-specific neutral species. The neutral species are then guided into the afterglow reaction tube for ionization via reactions with plasma-generated ions. For example, we recently demonstrated that concurrent introduction of organochlorines and methanolic sodium acetate into the ICP results in formation of NaCl, which

- is then ionized in the afterglow via adduction to formate ion to produce [NaClHCO2] [23]. Our efforts, however, did not yield detectable ions in negative mode for fluorine analysis. Importantly,

125 we reported that the difficulty in fluorine detection was likely related to the low formate affinity of NaF [23].

Here, we report a new ionization route to detect F by afterglow ionization of NaF via

Na+ adduction. Our experiments are conducted using aqueous solutions mixed with 10-mM NaOH in an online fashion prior to introduction into the ICP to enhance NaF formation. To investigate the available positive mode ionization mechanisms in the ICP afterglow, we examined the background ions in our experimental conditions. Figure 4.2 depicts dominance of Na+ and its

+ + adduct ions (mainly Na2OH ) in the background spectra, highlighting Na adduction as the major ionization pathway for neutrals in the ICP afterglow. Note that similar background spectra are observed using sodium acetate instead of NaOH (Figure 4.3) in pre-ICP mixing, indicating that

+ Na2OH is a result of plasma reactions rather than NaOH introduction.

Table 4.1. Reaction thermochemistries (kJ/mol) calculated at the ωB97xD/aug-cc-pVTZ level Reaction ΔH600K ΔG600K + + 1. Na + NaF  Na2F -246 -187 + + 2. Na + NaOH  Na2OH -235 -179 + + 3. Na + NaO2  [Na2O2] -216 -160

+ + The Na adduction mechanism opens the possibility of ionizing NaF into Na2F . This possibility is further strengthened by ab initio calculations (Table 4.1), revealing higher Na+ affinity for NaF relative to NaOH and NaO2 (neutrals for major ions in Figure 4.2). Therefore, NaF formed upon injection of fluorinated compounds into the plasma is expected to either adduct

+ + + directly with Na or undergo Na transfer with background ions in the afterglow, yielding Na2F .

Sequential flow injections of 50 μM p-fluoro-phenylalanine (p-F-phe) and 50 μM NaF with online

+ NaOH mixing shown in Figure 4.4 experimentally confirmed the formation of Na2F , offering a new ionization pathway for elemental detection of F from fluorochemicals.

126

Figure 4.2. Background spectrum of PARCI-MS with 80 μl/min 10-mM aqueous NaOH mixed with 250 μL/min HPLC water prior to introduction into the nebulizer.

127

Figure 4.3. Background spectrum of PARCI-MS using aqueous 10 mM sodium acetate for post-flow injection (post-column) sodium addition.

128

Figure 4.4. Flow injections of 50 µM NaF (first triplicate) and 50 µM p-F-phe (second

+ triplicate) lead to formation of Na2F in PARCI-MS. Note that difference in peak heights for the two compounds is a result of differences in flow injection peak shapes.

129

The fundamental difference between ionizations in PARCI-MS and conventional ICP-

MS/MS for F detection is evident from the experimental setups depicted in Figure 4.1, even though both techniques utilize ICP. In the conventional ICP-MS/MS, BaF+ is formed inside the ICP upon concurrent introduction of Ba and fluorinated compounds into the plasma. Reactions of ions with oxygen within a collision cell are then used to separate the analytical ion (BaF+) from isobaric interferences (e.g. Ba18OH+) [16–19]. In contrast, ion formation in PARCI occurs via afterglow ionization of plasma-generated analytical neutrals. This fundamental difference leads to major advantages for PARCI as discussed below.

4.4. Cool Plasma Operation

Figure 4.5a compares the effect of aerosol gas flow rate on sensitivity (ion current per

+ concentration unit of F) in PARCI-MS and conventional ICP-MS/MS. Notably, Na2F sensitivity in PARCI-MS optimizes at much higher aerosol gas flow rates compared to that for BaF+ in conventional ICP-MS/MS, denoting cool plasma temperatures in PARCI. Further, a relatively

+ wide range of aerosol gas flow rates (2.0-2.4 L/min) produce similar Na2F intensities in PARCI-

MS while a much narrower optimum range (1.3-1.35 L/min) is observed for BaF+ signal in ICP-

MS/MS. These observations indicate that a precise control of plasma temperature by the aerosol gas flow rate is needed for efficient formation and thermal ionization of BaF at the ion sampling

+ point within the ICP. On the other hand, Na2F formation shows less susceptibility to aerosol gas flow rate, suggesting resilience to plasma cooling and robustness upon plasma temperature fluctuations.

130

+ BaF using ICP-MS/MS

+ Na2F using PARCI-MS

10 µg/g F BaF+ using ICP- MS/MS 0.95 µg/g F

+ Na2F using PARCI-MS

Figure 4.5. Effect of aerosol carrier gas flow rate in ICP-MS/MS (BaF+) and PARCI-MS

+ (Na2F ) on a) absolute sensitivities and b) signal-to-background ratios corrected to 1000 ppb

F. Data for BaF+ are taken from reference 19. Error bars for PARCI-MS represent 95% confidence intervals based on triplicate flow injections of 50 μM p-F-phe (950 ppb).

131

4.5. Sensitivity

The overall sensitivity is dependent upon ion formation efficiency as well as the ion sampling and ion transmission efficiencies of the MS. Figure 4.5a illustrates that about 2 orders of magnitude better sensitivity is observed in PARCI-MS relative to ICP-MS/MS. We have measured

+ sensitivities of 180-500 cps/ppb for F using Na2F in PARCI-MS (depending on position of the reaction tube relative to the MS inlet) while values of 1.6-3.2 cps/ppb [16, 18, 19] and 0.026 cps/ppb [14] are reported for BaF+ using ICP-MS/MS and F+ in ICP-HR-MS, respectively.

Notably, the detected ion intensity in conventional ICP-MS/MS increases as oxygen is introduced into the reaction cell because of increased ion transmission via collisional focusing of BaF+ ions

[16]. This indicates efficient ion transmission through the mass analyzers. Therefore, the dramatically higher ion intensity in PARCI-MS is attributed to enhanced ionization and ion sampling efficiencies. The large flux of Ar+ ions in conventional ICP-MS severely reduces ion transmission from the ion source into the MS [27]. In contrast, cool plasma and atmospheric reaction tube in PARCI eliminate Ar+ ions while the high Na+ affinity of NaF ensures efficient formation and preservation of analytical ions in the afterglow, en route to the ion sampling point by the MS.

4.6. Signal-to-Background (S/B) Ratio

The S/B ratios for conventional ICP-MS/MS in Figure 4.5b closely mirror the trend for absolute BaF+ signal in Figure 4.5a, revealing wide variations in formation of BaF+ relative to the major isobaric interference Ba18OH+. In contrast, the S/B ratio for PARCI in Figure 4.5b remain

+ constant, despite variations in absolute Na2F signal intensity, illustrating that signal and background are affected similarly at various aerosol gas flow rates. The background at m/z 65 in

+ PARCI-MS is mainly composed of two species, 1) Na2F from fluorinated contaminants, and 2)

132

18 + + 17 + Na2 OH . Using isotopic distribution for Na2OH species and abundance of Na2 OH , we

18 + estimate that ~70% of the ion signal at m/z 65 is related to Na2 OH . This isobaric interference tempers the detection of low levels of F despite the high sensitivity in PARCI-MS. To estimate detection limit, ten injections of 10 μM NaF were conducted (Figure 4.6) and the LOD was calculated using 3×σheight×10/average peak height, yielding an LOD of 2.4 μM (46 ng/g) F in aqueous solutions. This is similar to LODs of 20-60 ng/g reported for ICP-MS/MS methods [16,

18, 19]. However, one must note that the analytical performance of ICP-MS/MS is achieved by reduction of isobaric interferences while no such measures have been taken in our studies.

Importantly, BaF+ is undetectable at 2500 ng/g level in ICP-MS without the use of MS/MS [16], further highlighting the fundamental advantage of chemical ionization in our approach. Moreover,

+ Na2F ions are thermodynamically stable in low temperature and atmospheric pressure and are readily detected by atmospheric-sampling mass spectrometers as shown in our experiments.

Accordingly, isobaric interferences can be resolved by facile coupling of PARCI to MS/MS and high-resolution molecular MS instruments without modification of the MS interface. The compatibility of PARCI with existing molecular MS platforms also enhances the accessibility of elemental quantitation, while the conventional ICP-MS approaches require dedicated elemental analysis instruments.

4.7. Compound-Independent Response

The relative peak areas from 50 μM flow injections of NaF and p-F-phe shown in Figure 4.7 illustrate that F detection is achieved with compound-independent sensitivities across a wide range of aerosol gas flow rates (2.0-2.5 L/min). Accordingly, we infer that formation of NaF in the plasma occurs in a robust manner over a wide range of plasma properties.

133

Figure 4.6. Ten replicate flow injections of 10 µM NaF used for estimation of limit of detection.

134

Figure 4.7. Effect of aerosol carrier gas flow rate on ratio of flow injection peak areas for 50

μM injections of sodium fluoride and p-F-phe in PARCI-MS. Error bars represent 95% confidence intervals based on triplicate flow injections.

135

4.8. Quantitation in Matrix

To illustrate the applications of PARCI in F analysis, we examined quantitation of fluoride in infant formula. The presence of fluoride in infant formula can cause fluorosis of infant teeth [28].

Therefore, methods for F- quantitation in infant formula are needed for food quality assurance [29].

This is a challenging analytical task because of the complex matrix of infant formula and difficulties for separation of fluoride from the matrix. The high matrix tolerance and specificity of

F-elemental analysis can address this challenge. We examined this potential by coupling ion chromatography (IC) to PARCI-MS. The IC eluent was split and mixed with NaOH prior to introduction into PARCI-MS. Importantly, sample preparation merely consisted of centrifugal filtering of the sample to separate high-molecular weight fats and proteins prior to injection into

IC-PARCI-MS. Details for operation of IC-PARCI-MS and sample preparation are provided in the SI.

+ - Figure 4.8a shows the Na2F ion signal for blank and 15-μM F spiked infant formula samples as well as a fluoride standard in water, indicating facile detection of fluoride in a complex matrix using IC-PARCI-MS. Figure 4.8b depicts results of similar experiments using ion chromatography with the conductivity detector. Note that PARCI-MS is placed in series with the conductivity detector, resulting in a slight delay between the retention times recorded by the two detectors for the same separation. Importantly, F- in aqueous standards is readily detected using conductivity detector (Figure 4.8b), however, matrix in infant formula completely masks the signal for F-, rendering this detector unusable in the presence of matrix. Interestingly, the shape of the IC-

PARCI-MS peak for spiked infant formula in Figure 4.8a is similar to that of the matrix in Figure

4.8b, while F- elutes in a sharp peak for standards prepared in water (Figures 4.8a and 4.8b). This indicates that the packing of F- on the IC column is severely impacted by the presence of matrix

136 in addition to its co-elution with the matrix components. Nonetheless, an average recovery of 83% and repeatability of 6.3% RSD are achieved for triplicate analysis of the infant formula by IC-

PARCI-MS at a spike level of 15 μM (285 ng/g) using a calibration curve with F- standards prepared in water (Figure 4.9). These results highlight the matrix tolerance and quantitative advantage of elemental mass spectrometry using PARCI. Notably, the spike level in Figure 4.8a is below the quantification limit required by a recent call for fluoride detection methods in infant formula (300 ng/g) [29], denoting good sensitivity of this technique. Similar investigation of 52

μM (988 ng/g) spike level yielded average recovery of 98% and repeatability of 3.5 RSD%.

In conclusion, we present proof of principles that afterglow Na+ adduction of ICP-generated

NaF from fluorinated compounds offers an efficient and robust ionization route for elemental F detection and quantitation. This approach is in contrast to those used in conventional ICP-MS where ionization occurs inside the plasma. Consequently, fundamental challenges in ion formation and ion sampling with conventional ICP-MS are circumvented, yielding over two orders of magnitude better sensitivity for F. Further, similar response factors from inorganic and organic F- containing compounds are obtained over a wide range of aerosol gas flow rates, highlighting the elemental nature of ionization and offering compound-independent quantitation. We have shown the capabilities of this ionization approach in analysis of infant formula as an example of a complex sample, demonstrating high matrix tolerance. Notably, afterglow ionization renders this approach readily implementable on molecular mass spectrometry platforms where improved performance could be obtained by elimination of isobaric interferences via MS/MS and high-resolution MS.

Overall, the presented technique provides an approach to enhance discovery and quantitation of fluorinated compounds of rising importance in pharmaceutical, environmental, and food quality analyses.

137

- - a 15 μM F 6.7 μM F in 1.9 spiked infant water

formula 1.7 blank infant 1.5

cps) formula

intensity intensity

5

+

F (10

2 1.3 Na 1.1

0.9

b - 15 μM F

blank infant 80 spiked infant S) formula μ formula - 9 μM F in

40 water

Conductivity ( Conductivity 0 6 6.5 7 7.5 8 8.5 9 9.5

Time (min)

Figure 4.8. Chromatograms for blank and spiked infant formula samples as well as fluoride standards in water using a) IC-PARCI-MS and b) IC-conductivity detector.

138

Figure 4.9. Calibration curve for IC-PARCI-MS using NaF standards in water with a linear least-squares fit.

139

4.9. References

1. Fujiwara, T., O’Hagan, D.: Successful fluorine-containing herbicide agrochemicals. J.

Fluor. Chem. 167, 16–29 (2014). https://doi.org/10.1016/j.jfluchem.2014.06.014

2. O’Hagan, D.: Fluorine in health care: Organofluorine containing blockbuster drugs. J.

Fluor. Chem. 131, 1071–1081 (2010). https://doi.org/10.1016/j.jfluchem.2010.03.003

3. Houtz, E.F., Higgins, C.P., Field, J.A., Sedlak, D.L.: Persistence of Perfluoroalkyl Acid

Precursors in AFFF-Impacted Groundwater and Soil. Environ. Sci. Technol. 47, 8187–8195

(2013). https://doi.org/10.1021/es4018877

4. Li, S., Huang, J., Chen, Z., Chen, G., Lai, Y.: A review on special wettability textiles: theoretical models, fabrication technologies and multifunctional applications. J. Mater. Chem. A.

5, 31–55 (2017). https://doi.org/10.1039/C6TA07984A

5. Schaider, L.A., Balan, S.A., Blum, A., Andrews, D.Q., Strynar, M.J., Dickinson, M.E.,

Lunderberg, D.M., Lang, J.R., Peaslee, G.F.: Fluorinated Compounds in U.S. Fast Food

Packaging. Environ. Sci. Technol. Lett. 4, 105–111 (2017). https://doi.org/10.1021/acs.estlett.6b00435

6. Harding-Marjanovic, K.C., Houtz, E.F., Yi, S., Field, J.A., Sedlak, D.L., Alvarez-Cohen,

L.: Aerobic Biotransformation of Fluorotelomer Thioether Amido Sulfonate (Lodyne) in AFFF-

Amended Microcosms. Environ. Sci. Technol. 49, 7666–7674 (2015). https://doi.org/10.1021/acs.est.5b01219

7. Yeung, L.W.Y., Mabury, S.A.: Are humans exposed to increasing amounts of unidentified organofluorine? Environ. Chem. 13, 102–110 (2015). https://doi.org/10.1071/EN15041

8. Gong, Y., Chen, J., Shi, Y., Lim, H.-K., Weng, N., Salter, R.: Standard-Free Bioanalytical

Approach for Absolute Quantitation of Drug Metabolites Utilizing Biosynthesis of Reciprocal

140

Radio and Stable Isotopologues and Its Application. Anal. Chem. 89, 8399–8404 (2017). https://doi.org/10.1021/acs.analchem.7b01830

9. McDonough, C.A., Guelfo, J.L., Higgins, C.P.: Measuring total PFASs in water: The tradeoff between selectivity and inclusivity. Curr. Opin. Environ. Sci. Health. 7, 13–18 (2019). https://doi.org/10.1016/j.coesh.2018.08.005

10. O’Hagan, D., Deng, H.: Enzymatic Fluorination and Biotechnological Developments of the Fluorinase. Chem. Rev. 115, 634–649 (2015). https://doi.org/10.1021/cr500209t

11. Jamari, N.L.A., Dohmann, J.F., Raab, A., Krupp, E.M., Feldmann, J.: Novel non-targeted analysis of perfluorinated compounds using fluorine-specific detection regardless of their ionisability (HPLC-ICPMS/MS-ESI-MS). Anal. Chim. Acta. 1053, 22–31 (2019). https://doi.org/10.1016/j.aca.2018.11.037

12. Klencsár, B., Li, S., Balcaen, L., Vanhaecke, F.: High-performance liquid chromatography coupled to inductively coupled plasma – Mass spectrometry (HPLC-ICP-MS) for quantitative metabolite profiling of non-metal drugs. TrAC Trends Anal. Chem. 104, 118–134 (2018). https://doi.org/10.1016/j.trac.2017.09.020

13. Calderón-Celis, F., Sanz-Medel, A., Ruiz Encinar, J.: Universal absolute quantification of biomolecules using element mass spectrometry and generic standards. Chem. Commun. 54, 904–

907 (2018). https://doi.org/10.1039/C7CC09059E

14. Bu, X., Wang, T., Hall, G.: Determination of halogens in organic compounds by high resolution inductively coupled plasma mass spectrometry (HR-ICP-MS). J. Anal. At. Spectrom.

18, 1443–1451 (2003). https://doi.org/10.1039/B306570G

141

15. Vickers, G.H., Wilson, D.A., Hieftje, G.M.: Detection of negative ions by inductively coupled plasma mass spectrometry. Anal. Chem. 60, 1808–1812 (1988). https://doi.org/10.1021/ac00168a031

16. Jamari, N.L.A., Dohmann, J.F., Raab, A., Krupp, E.M., Feldmann, J.: Novel non-target analysis of fluorine compounds using ICPMS/MS and HPLC-ICPMS/MS. J. Anal. At. Spectrom.

32, 942–950 (2017). https://doi.org/10.1039/C7JA00051K

17. Jamari, N.L.A., Behrens, A., Raab, A., Krupp, E.M., Feldmann, J.: Plasma processes to detect fluorine with ICPMS/MS as [M–F]+: an argument for building a negative mode

ICPMS/MS. J. Anal. At. Spectrom. 33, 1304–1309 (2018). https://doi.org/10.1039/C8JA00050F

18. Guo, W., Jin, L., Hu, S., Guo, Q.: Method Development for the Determination of Total

Fluorine in Foods by Tandem Inductively Coupled Plasma Mass Spectrometry with a Mass-Shift

Strategy. J. Agric. Food Chem. 65, 3406–3412 (2017). https://doi.org/10.1021/acs.jafc.7b00535

19. Zhu, Y., Nakano, K., Shikamori, Y.: Analysis of Fluorine in Drinking Water by ICP-

QMS/QMS with an Octupole Reaction Cell. Anal. Sci. 33, 1279–1284 (2017). https://doi.org/10.2116/analsci.33.1279

20. Qin, Z., McNee, D., Gleisner, H., Raab, A., Kyeremeh, K., Jaspars, M., Krupp, E., Deng,

H., Feldmann, J.: Fluorine Speciation Analysis Using Reverse Phase Liquid Chromatography

Coupled Off-Line to Continuum Source Molecular Absorption Spectrometry (CS-MAS):

Identification and Quantification of Novel Fluorinated Organic Compounds in Environmental and

Biological Samples. Anal. Chem. 84, 6213–6219 (2012). https://doi.org/10.1021/ac301201y

21. Qin, Z., Raab, A., Krupp, E., Deng, H., Feldmann, J.: Mining complex bacteria media for all fluorinated compounds made possible by using HPLC coupled parallel to fluorine-specific and

142 molecular specific detection. J. Anal. At. Spectrom. 28, 877–882 (2013). https://doi.org/10.1039/C3JA50086A

22. Lesniewski, J.E., McMahon, W.P., Zheng, K., Wang, H., Badiei, H., Jorabchi, K.:

Atmospheric pressure plasma assisted reaction chemical ionization for analysis of chlorinated compounds separated by liquid chromatography. J. Anal. At. Spectrom. 32, 1757–1765 (2017). https://doi.org/10.1039/C7JA00115K

23. Lesniewski, J.E., McMahon, W.P., Jorabchi, K.: Mechanistic insights into chloride ion detection from the atmospheric-pressure afterglow of an argon inductively coupled plasma. J Anal

Spectrom. 33, 1981–1992 (2018). https://doi.org/10.1039/c8ja00244d

24. Frisch, M.J., Trucks, G.W., Schlegel, H.B., Scuseria, G.E., Robb, M.A., Cheeseman, J.R.,

Scalmani, G., Barone, V., Petersson, G.A., Nakatsuji, H., Li, X., Caricato, M., Marenich, A.V.,

Bloino, J., Janesko, B.G., Gomperts, R., Mennucci, B., Hratchian, H.P., Ortiz, J.V., Izmaylov,

A.F., Sonnenberg, J.L., Williams, Ding, F., Lipparini, F., Egidi, F., Goings, J., Peng, B., Petrone,

A., Henderson, T., Ranasinghe, D., Zakrzewski, V.G., Gao, J., Rega, N., Zheng, G., Liang, W.,

Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y.,

Kitao, O., Nakai, H., Vreven, T., Throssell, K., Montgomery Jr., J.A., Peralta, J.E., Ogliaro, F.,

Bearpark, M.J., Heyd, J.J., Brothers, E.N., Kudin, K.N., Staroverov, V.N., Keith, T.A., Kobayashi,

R., Normand, J., Raghavachari, K., Rendell, A.P., Burant, J.C., Iyengar, S.S., Tomasi, J., Cossi,

M., Millam, J.M., Klene, M., Adamo, C., Cammi, R., Ochterski, J.W., Martin, R.L., Morokuma,

K., Farkas, O., Foresman, J.B., Fox, D.J.: Gaussian 16 Rev. B.01. , Wallingford, CT (2016)

25. Hyttinen, N., Otkjær, R.V., Iyer, S., Kjaergaard, H.G., Rissanen, M.P., Wennberg, P.O.,

Kurtén, T.: Computational Comparison of Different Reagent Ions in the Chemical Ionization of

143

Oxidized Multifunctional Compounds. J. Phys. Chem. A. 122, 269–279 (2018). https://doi.org/10.1021/acs.jpca.7b10015

26. Hanwell, M.D., Curtis, D.E., Lonie, D.C., Vandermeersch, T., Zurek, E., Hutchison, G.R.:

Avogadro: an advanced semantic chemical editor, visualization, and analysis platform. J.

Cheminformatics. 4, 17 (2012). https://doi.org/10.1186/1758-2946-4-17

27. Egorova, T., Dietiker, R., Hattendorf, B., Günther, D.: Characteristics of an ion funnel as ion guide in an inductively coupled plasma mass spectrometer. Spectrochim. Acta Part B At.

Spectrosc. 76, 40–47 (2012). https://doi.org/10.1016/j.sab.2012.06.010

28. Berg, J., Gerweck, C., Hujoel, P.P., King, R., Krol, D.M., Kumar, J., Levy, S., Pollick, H.,

Whitford, G.M., Strock, S., Aravamudhan, K., Frantsve-Hawley, J., Meyer, D.M., for the

American Dental Association Council on Scientific Affairs Expert Panel on Fluoride Intake From

Infant Formula and Fluorosis: Evidence-Based Clinical Recommendations Regarding Fluoride

Intake From Reconstituted Infant Formula and Enamel Fluorosis: A Report of the American Dental

Association Council on Scientific Affairs. J. Am. Dent. Assoc. 142, 79–87 (2011). https://doi.org/10.14219/jada.archive.2011.0032

29. AOAC SMPR® 2014.016, http://www.eoma.aoac.org/SMPR/upload/45/SMPR%202014_016.pdf

144

Chapter V§

Summary and Future Directions

5.1. Summary of Core Findings

The findings reported in this dissertation demonstrate the analytical utility and flexibility of the plasma-assisted reaction chemical ionization (PARCI) approach for ionization and quantitation of nonmetals. Specifically, this approach offers both fundamental and practical advantages when compared with the current state-of-the-art elemental analysis method, ICP-MS.

The use of chemical, rather than thermal ionization allows the fundamental ionization potential barriers of nonmetal elements to be overcome. The use of an atmospheric-pressure reaction interface permits efficient ion-neutral reactions to occur in the afterglow of an inductively coupled plasma and allows for facile coupling to mass spectrometers designed for molecular ionization sources.

In Chapter II we described a new atmospheric-pressure reaction interface that allows plasma gasses to cool for improved chemical ionization of nonmetals, in particular Cl. We demonstrated compound-independent response factors and improved sensitivity for chlorine analysis. The mechanism of ionization utilized in Chapter II was explored in Chapter III.

Importantly, the roles of sodium and methanol in chloride ion formation were clarified; fundamental investigations revealed the formation of an intermediate species, namely NaCl in gas phase. However, the same ionization mechanism was ineffective for F detection due to less favorable reaction energetics for NaF. Chapter IV demonstrated a new ionization chemistry to

+ overcome this limitation by converting NaF neutrals to the thermodynamically favorable Na2F .

Notably, we observed two-orders-of-magnitude-improved sensitivity for F detection compared to

§ The material in 5.2.1 is based upon work supported by the National Science Foundation (NSF) under CHE- 1507304. We would also like to thank Georgetown University for financial support of the work in 5.2.2. 145

ICP-MS/MS methods while similar detection limits were obtained. Improved sensitivities for chlorine and fluorine demonstrate that we have successfully achieved our goal of overcoming the fundamental limitations of ICP-MS for ionization of these elements. In the next section of this chapter, we will discuss preliminary data and future directions for the development of the work presented in this dissertation.

5.2. Opportunities for Expansion of this Work

The general approach described in this dissertation of employing chemical ionization to convert neutral plasma products to ions can be applied to other nonmetal elements. Previous chapters have outlined the importance of fluorine and chlorine in the environment and pharmaceuticals, however phosphorous and sulfur are also of particular interest due to their biological relevance [1]. In order to maximize the applicability of PARCI to biological samples, analytical ions for both P and S need to be identified.

Moreover, the tunability of the chemical ionization in PARCI may also help alleviate another challenge in ICP-MS: simultaneous quantitation of nonmetals. Simultaneous quantitation of multiple nonmetals in ICP-MS is fundamentally limited due to the formation of analytical ions for F, Cl, P, and S in different plasma conditions. For example, the use of cool plasma conditions in ICP-MS/MS for BaF+ detection is required which would reduce the ionization efficiency of Cl+,

P+, and S+. Furthermore, the limits of detection for chlorine would be compromised by the use of

+ + + an O2 collision gas that is needed for BaF [2–5], P , and S [6–8] detection in place of the H2 collision gas that is typically used to reduce the isobaric interferences of Cl+ [9].

In the PARCI methods described in the preceding chapters, chlorine is detected as Cl- while

+ Na2F is used for fluorine with different instrument operating parameters. Consequently, in order to provide simultaneous quantification of F, Cl, P, and S we will need to identify ions that form

146 efficiently for all four of these elements in similar operating conditions. Preliminary data are presented below and briefly examined for the detection of phosphorous and sulfur as well as chlorine, offering the possibility of further development toward efficient simultaneous detection of all four elements.

5.2.1. Negative-Mode Phosphorous Detection using PARCI

- Phosphorous can be detected in negative-mode as PO3 , as demonstrated in Figure 5.1a.

Experimental conditions used for the data in Figure 5.1 are summarized in Table 5.1. The formation of this polyatomic ion is not completely unexpected given that it has also been reported in negative-mode ICP-MS [10, 11]. The formation mechanisms for this ion, however, are currently

- unknown in both systems. A heat-map of PO3 flow injection peak area for 1 µM Na2HPO4 in H2O

(as a measure of sensitivity) at carrier gas flow rates between 1.8-2.8 SLPM and exhaust velocities from 3.8-6.8 m/s is seen in Figure 5.1b. The carrier gas flow rate appears to have a more significant effect than exhaust velocity, suggesting that the plasma temperature is more critical compared to reaction tube temperature for formation of this ion.

- Table 5.1. Operating parameters for PO3 experiments Parameter Value Plasma gas 14 L/min Auxiliary gas 1.2 L/min RF power 1100W (700-1500W for power effect) Injector tip diameter 1.5 mm Analyte nebulizer gas 1.25 SLPM Reagent nebulizer gas 0.55 SLPM Makeup gas 0-1 SLPM (0.3 SLPM for power effect) Analyte solution flow rate 200 µL/min LC-MS water Reagent solution 200 µM sodium acetate in LC-MS grade methanol Reagent solution flow rate 30 µL/min Exhaust velocity 3.8, 4.8, 5.8, 6.8 m/s (5.8 m/s power effect) Sampler Flat aluminum orifice, 2 mm id, 1 mm thick Sampling depth 9 mm downstream of the load coil Reaction tube 50 mm stainless steel tube, ¼” o.d., 0.19” i.d Injection loop 20 µL stainless steel

147

a b

Peak Area

-

3

PO

Intensity (CPS) Intensity

-

3 PO

Time (min) 3E+06 c 2E+06

2E+06

Peak Area Peak 1E+06

- 3

5E+05 PO

0E+00 650 850 1050 1250 1450 1650 RF Power (W)

- Figure 5.1. (a) Typical flow injections of 1 µM Na2HPO4 in H2O monitoring PO3 (m/z 79).

- (b) a heatmap showing PO3 (m/z 79) peak areas for flow injections of 1 µM Na2HPO4 in H2O as a function of carrier flow rate and exhaust velocity. (c) the effect of power on peak areas for flow injections of 1 µM Na2HPO4 in H2O.

148

- The effect of RF power on sensitivity of PO3 in the middle of the optimum operating zone

- (5.8 m/s exhaust velocity, 2.1 L/min carrier gas in Figure 5.1b) is detailed in Figure 5.1c. PO3 optimizes at 1100W, highlighting the need for cool plasma conditions for formation of this ion.

Preliminary sensitivity and limits of detection using 5.8 m/s exhaust velocity and 2.2 L/min carrier gas flow rate, are estimated at 5950 cps/ppb and 2 ppb, respectively – inferior to the ppt range detection limits reported for positive-mode ICP-MS/MS methods [6–8, 12]. Furthermore,

- detection limits for PO3 are currently background limited, indicating that our instrument has a phosphorous contamination or abundant isobaric interferences. Elimination of the background would significantly improve detection limits; however, sensitivity improvements are also needed in order to exceed the performance of positive-mode ICP-MS/MS. These analytical figures of merit are however, approximately two orders of magnitude better than those reported for the same

- ion (PO3 ) in negative-mode ICP-MS, which offers 0.3 cps/ppb sensitivity with 200 ppb limits of detection [10], indicating that this ion likely forms as plasma gasses cool. Importantly, the broad sensitivity optimum observed in Figure 5.1b includes typical chlorine operating conditions

(1100W, 2.0 L/min carrier, 5.8 m/s exhaust), suggesting that simultaneous quantitation of

- - phosphorous as PO3 and chlorine as Cl is possible.

5.2.2. Positive-Mode Phosphorous, Sulfur, and Chlorine Detection using PARCI

In order to enable simultaneous quantitation of F, Cl, P and S it is preferable to ionize all of these elements in a single polarity. Since we were not able to identify stable fluorinated analytical ions in negative-mode, we investigated positive mode ions for P, S, and Cl detection.

Figures 5.2a, 5.3a, and 5.4a show flow injections of phosphorous-, sulfur-, and chlorine-containing

+ + + compounds and their detection as Na3HPO4 , Na3SO4 , and Na2Cl , respectively in the same operating conditions summarized in Table 5.2. Figures 5.2b, 5.3b, and 5.4b show compound-

149 independent response for each of these elements, where response factors, defined as the flow injection peak area divided by the elemental concentration injected, are normalized to the average response factor of all compounds on the given graph.

Table 5.2. Operating parameters for simultaneous monitoring of P, S, Cl, and F in positive-mode Parameter Value Plasma gas 14 L/min Auxiliary gas 1.2 L/min RF power 1100W Injector tip diameter 1.5 mm Analyte nebulizer gas 1.1 SLPM Reagent nebulizer gas 1.0 SLPM Makeup gas 0 Analyte solution flow rate 200 µL/min 50% water:methanol Reagent solution 10 mM sodium acetate in 50% water:methanol Reagent solution flow rate 80 µL/min Exhaust velocity 4.8 m/s Sampler Flat aluminum orifice, 2 mm id, 1 mm thick Sampling depth 9 mm downstream of the load coil Reaction tube 50 mm stainless steel tube, ¼” o.d., 0.19” i.d Injection loop 20 µL stainless steel

Estimated limits of detection and sensitivities for these three ions are described in Table

+ 5.3. Na2F performance in the same operating conditions is also included as a proof-of-concept for simultaneous detection of all four elements using PARCI, a feat that is fundamentally challenging using ICP-MS/MS due to ion formation in different operating conditions. Sensitivities and detection limits for ICP-MS/MS optimized separately for measurement of each element are given in Table 5.4 for comparison. At first glance, sensitivities and detection limits for chlorine using

+ + Na2Cl appear worse than those of Cl methods in ICP-MS/MS [9], while Fluorine detection is still significantly more sensitive than BaF+ methods on ICP-MS/MS [2–5, 13]. These comparisons greatly underestimate the performance of PARCI however, since ICP-MS/MS was optimized for best performance for measuring a single element.

150

1 µM glyphosate 1 µM Na2HPO4 a

Intensity (CPS) Intensity

+

4

O

HP

3 Na

Time (min)

1.8 b 1.6 Na2HPO4 1.4 glyphosate 1.2

1

0.8

0.6

0.4

Normalized Response Factor Response Normalized 0.2

0 Compound

Figure 5.2. (a) Triplicate flow injections of 1 µM glyphosate in 50% methanol:H2O followed

+ by triplicate injections of 1 µM Na2HPO4 in 50% methanol:H2O monitoring Na3HPO4 (m/z

165). (b) Normalized response factor plot showing compound-independent response for

+ phosphorous detected as Na3HPO4 (m/z 165).

151

50 µM Na2SO4 a

50µM methionine

Intensity (CPS) Intensity

+

4

SO

3 Na

Time (min)

1.8 b 1.6 1.4 Na SO 2 4 methionine 1.2

1

0.8

0.6

0.4

Normalized Response Factor Response Normalized 0.2

0 Compound

Figure 5.3. (a) Triplicate flow injections of 50 µM Na2SO4 and 50 µM methionine in 50%

+ methanol:H2O monitoring Na3SO4 (m/z 165). (b) Normalized response factor plot showing

+ compound-independent response for sulfur detected as Na3SO4 (m/z 165).

152

25 µM Chloramphenicol 50 µM NaCl a

Intensity (CPS) Intensity

+

Cl

35

2 Na

Time (min)

1.8 b Chloramphenicol 1.6 NaCl 1.4

1.2

1

0.8

0.6

0.4

Normalized Response Factor Response Normalized 0.2

0 Compound

Figure 5.4. (a) Triplicate flow injections of 25 µM chloramphenicol (50 µM in Cl) in H2O and

35 + 50 µM NaCl in H2O monitoring Na2 Cl (m/z 81). (b) Normalized response factor plot showing

+ compound-independent response for chlorinated compounds detected as Na2Cl .

153

Notably, detection of analytical ions for F, Cl, P, and S, denotes the potential of PARCI for enhanced simultaneous quantification of the four elements. Further, facile coupling of the ionization method to many different types of mass spectrometers help to push the boundaries of sensitivity and detection limits for these elements.

Table 5.3. Preliminary estimates of sensitivity and detection limits for simultaneous detection of P, S, Cl and F analytical ions using PARCI Element Ion Sensitivity (cps/ppb) LOD (ppb) + P Na3HPO4 1157 0.51 + S Na3SO4 46 19 + Cl Na2Cl 188 12 + F Na2F 103 44

Table 5.4. Sensitivities and detection limits for ICP-MS/MS optimized for detection of a single element Sensitivity Element Ion LOD (ppb) Reference (cps/ppb) + + a P P as PO2 22 0.016 7 S S+ as SO+ 6821 0.1 7a + b Cl Cl as ClH2+ 287 1 9 F BaF+ 2 60 5c a P and S optimize in the same conditions, solvent is 20% methanol:H2O bSolvent is methanol c Solvent is H2O

5.3. References

1. Calderón-Celis, F., Sanz-Medel, A., Ruiz Encinar, J.: Universal absolute quantification of biomolecules using element mass spectrometry and generic standards. Chemical Communications.

54, 904–907 (2018). https://doi.org/10.1039/C7CC09059E

2. Jamari, N.L.A., Dohmann, J.F., Raab, A., Krupp, E.M., Feldmann, J.: Novel non-target analysis of fluorine compounds using ICPMS/MS and HPLC-ICPMS/MS. J. Anal. At. Spectrom.

32, 942–950 (2017). https://doi.org/10.1039/C7JA00051K

154

3. Jamari, N.L.A., Behrens, A., Raab, A., M. Krupp, E., Feldmann, J.: Plasma processes to detect fluorine with ICPMS/MS as [M–F] + : an argument for building a negative mode

ICPMS/MS. Journal of Analytical Atomic Spectrometry. 33, 1304–1309 (2018). https://doi.org/10.1039/C8JA00050F

4. Jamari, N.L.A., Dohmann, J.F., Raab, A., Krupp, E.M., Feldmann, J.: Novel non-targeted analysis of perfluorinated compounds using fluorine-specific detection regardless of their ionisability (HPLC-ICPMS/MS-ESI-MS). Analytica Chimica Acta. 1053, 22–31 (2019). https://doi.org/10.1016/j.aca.2018.11.037

5. Zhu, Y., Nakano, K., Shikamori, Y.: Analysis of Fluorine in Drinking Water by ICP-

QMS/QMS with an Octupole Reaction Cell. Analytical Sciences. 33, 1279–1284 (2017). https://doi.org/10.2116/analsci.33.1279

6. Nakano, K.: Ultra-low level determination of phosphorus, sulfur, silicon and chlorine using the Agilent 8900 ICP-QQQ. In: Ultra-low level determination of phosphorus, sulfur, silicon and chlorine using the Agilent 8900 ICP-QQQ. pp. 1–5. Agilent Technologies (2016)

7. Soffey, E., Woods, B., Wilbur, S.: Analysis of silicon, phosphorus and sulfur in 20% methanol. In: AGILENT 8800 ICP-QQQ APPLICATION HANDBOOK. pp. 19–20. Agilent

Technologies (2015)

8. Sugiyama, N.: Analysis of sulfur, phosphorus, silicon and chlorine in N-methyl-2- pyrrolidone. In: AGILENT 8800 ICP-QQQ APPLICATION HANDBOOK. pp. 16–18. Agilent

Technologies (2015)

9. Klencsár, B., Bolea-Fernandez, E., Flórez, M.R., Balcaen, L., Cuyckens, F., Lynen, F.,

Vanhaecke, F.: Determination of the total drug-related chlorine and bromine contents in human blood plasma using high performance liquid chromatography–tandem ICP-mass spectrometry

155

(HPLC–ICP-MS/MS). Journal of Pharmaceutical and Biomedical Analysis. 124, 112–119 (2016). https://doi.org/10.1016/j.jpba.2016.02.019

10. Fulford, J.E., Quan, E.S.K.: Negative Ions in Inductively Coupled Plasma Mass

Spectrometry. Appl. Spectrosc., AS. 42, 425–428 (1988)

11. Vickers, G.H., Wilson, D.A., Hieftje, G.M.: Detection of negative ions by inductively coupled plasma mass spectrometry. Anal. Chem. 60, 1808–1812 (1988). https://doi.org/10.1021/ac00168a031

12. Fernández, S.D., Sugishama, N., Encinar, J.R., Sanz-Medel, A.: Triple Quad ICPMS

(ICPQQQ) as a New Tool for Absolute Quantitative Proteomics and Phosphoproteomics. Anal.

Chem. 84, 5851–5857 (2012). https://doi.org/10.1021/ac3009516

13. Guo, W., Jin, L., Hu, S., Guo, Q.: Method Development for the Determination of Total

Fluorine in Foods by Tandem Inductively Coupled Plasma Mass Spectrometry with a Mass-Shift

Strategy. J. Agric. Food Chem. (2017). https://doi.org/10.1021/acs.jafc.7b00535

156

Appendix

Permission Documents

Reuse Permission for Chapter II

157

Reuse Permission for Chapter III

158

Reuse Permission for Chapter IV

159