mm“ 9””76670 \|\\\\\\l|\\\\||\\llllllllllllllllll 3 1293 006 6050 LIBRARY A Michigan State University . This is to certify that the dissertation entitled Development of Sensitive Assays Based on Chemical Oxidation and Electron Capture Negative Ionization Mass Spectrometry for the Determination of Corticosteroids in Biological Fluids presented by Kathleen Ann Kayganich has been accepted towards fulfillment of the requirements for Ph . D . degree in Chemistry Major professor Date 8-1 1—89 MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. l—r: 1 —-—— DATE DUE DATE DUE DATE DUE “—7 MSU Is An Affirmative ActiotVEqual Opportunity Institution DEVELOPMENT OF SENSITIVE ASSAYS BASED ON CHEMICAL OXIDATION AND ELECTRON CAPTURE NEGATIVE IONIZATION MASS SPECTROMETRY FOR THE DETERMINATION OF CORTICOSTEROIDS IN BIOLOGICAL FLUIDS by Kathleen Ann Kayganich A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1989 CMOM‘I ABSTRACT DEVELOPMENT OF SENSITIVE ASSAYS BASED ON CHEMICAL OXIDATION AND ELECTRON CAPTURE NEGATIVE IONIZATION MASS SPECTROMETRY FOR THE DETERMINATION OF CORTICOSTEROIDS IN BIOLOGICAL FLUIDS By Kathleen Ann Kayganich The sensitivity and selectivity of electron capture negative ionization mass spectrometry (ECNI/MS) make this a valuable technique for the determination of low levels of analytes in complex matrices. However, most compounds of biological interest do not have a high electron capture response. Derivatization of functional groups with halogenated moieties is usually employed to enhance the electron capture responses of compounds having poor electron capture response. Often, derivatization causes an increase in the chemical background, so that the increase in the signal-to—background ratio for the analyte is not greatly enhanced. As an alternative to the use of derivatization, chemical oxidation has been explored as a more selective means to enhance electron capture response. The responses of some corticosteroid drugs are increased markedly upon oxidation. The responses of most endogenous steroids are not greatly enhanced by oxidation. Dexamethasone is a corticosteroid drug that is oxidized to a highly electrophilic analog. Sensitive methodology is needed to determine low levels of dexamethasone in human plasma. Procedures for the isolation and oxidation of dexamethasone were investigated to optimize the production and recovery of the oxidized analog. A sensitive assay was developed for dexamethasone in plasma using oxidation and ECNI-MS detection. The results obtained from the analyses of several plasma samples using this method were comparable to the results obtained by a conventional GC—MS method and by RIA in an interlaboratory study. The oxidation product of the major metabolite of dexamethasone, 6B- hydroxydexamethasone, was also detected by ECNI/MS in the analysis of plasma samples. Tandem mass spectrometry (MS/MS) was investigated for the detection of oxidized dexamethasone in the plasma matirx. The direct inlet probe (DIP) could be used to introduce the sample for analysis by ECNI/MS/MS due to the additional selectivity provided by this technique. An assay for 6B-hydroxycortisol, based on oxidation and ECNI/MS analysis, was also investigated. 6B-Hydroxycortisol is a minor component of urine that is useful as an index of microsomal enzyme activity. The results obtained for the determination of 6B- hydroxycortisol in human urine by this method were comparable to those obtained using a conventional GC-MS method. The oxidation method with GC-ECNI/MS analysis was more selective than the conventional GC-MS method. This allowed the use of less rigorous GC conditions and provided a shorter analysis. ACKNOWLEDGEMENTS I would like to acknowledge Doug Gage, Mike Davenport, Mel Micke, Melinda Beming, Betty Baltzer, and Bev Charnberlin, the staff of the Michigan State University Mass Spectrometry Facility, for the expert assistance and encouragement they have given me during my years at Michigan State. I especially appreciated their willingness to help at a moment’s notice. I would also like to thank the Watson group, past and present, for their help and encouragement. A thank-you to Linda Doherty, Dan Kassel and Brian Musselman for teaching me to boldly disassemble mass spectrometers and for teaching me many other things. Thanks especially to Tim Heath for helpful discussions and his willingness to work with me on the MS/MS experiments. I would also like to thank Mark McMills for all the valuable advice he gave me concerning organic chemistry and Blake Rees for the use of his computer. A special thanks to Curt Heine for his thoughtfulness and encouragement, and to Ellen Yurek, Helen Mayer, and Dave Wagner for their friendship and support. Thanks also to Ben Gardner for plumbing expertise and humor, to Gary Schultz for some good meals, to Karen Light for her thoughtfulness, for help with the HP5985, and for making ASMS hotel reservations year after year. I also want to thank my friends Lauren and Mark McMills, Deb de Vries, Liann Short, Pat Redmon, Helen Palmer, Dennis Rich, Mark and Sherre Whalon, Joan Stark, and my parents for listening and being their for me and helping me to grow. And, finally, I would like to thank Dr. Watson for his helpfulness - especially in preparation for ASMS presentations, for research support, for his encouragement and advice, the freedom he allowed me in the lab, and for salmon steaks and a good auto mechanic! iv TABLE OF CONTENTS List of Figures ........................................... viii List of Tables ............................................ xi Key to Abbreviations ....................................... xiii CHAPTER 1. INTRODUCTION ................................ 1 1. Mass Spectrometry in Biomedical Research ...................... 1 A. Advantages of Mass Spectrometry .......................... 1 1. Specificity ...................................... 1 2. Sensitivity ...................................... 4 B. Derivatization of Corticosteroids for Analysis by GC-MS ............ 4 1. Qualitative Gas Chromatography-Mass Spectrometry ............ 4 2. Quantitative Mass Spectrometry ......................... 5 3. Mass Spectrometry as a Reference Method .................. 6 H. Conventional Mass Spectrometry for Corticosteroids Analysis ........... 7 A. Corticosteroids ..................................... 8 B. Derivatization of Corticosteroids for Analysis by GC—MS ........... 10 C. Other Mass Spectrometric Techniques for Corticosteroids Analysis ..... 13 III. Electron Capture Negative Ion Mass Spectrometry .................. 14 A. History and Theory of Electron-Capture Negative Ionization ......... 14 B. Limitations of Electron Capture Negative Ionization .............. 15 C. Advantages of ECNI ................................. l7 1. Sensitivity ..................................... 17 2. Selectivity ..................................... 18 D. Derivatization to Enhance ECNI Response .................... 19 IV. An Alternative Approach for Enhancing ECNI Response of Corticosteroids . . .22 A. Previous work ..................................... 23 B. Majors goals of this research ............................ 26 References .............................................. 27 V CHAPTER 2. DEVELOPMENT OF OPTIMAL CONDITIONS FOR ISOLATION AND CHEMICAL OXIDATION OF DEXAMETHASONE ...... 30 I. Introduction ......................................... 3 0 11. Method Development ................................... 32 A. Extraction of Dexamethasone and 6B—Hydroxydexamethasone from Plasma ...................................... 32 1. Solvent Extraction vs. Solid Phase Extraction ..... , ........... 32 2. Solid-Phase, Reverse—Phase (C18) Extraction ................ 36 3. Influence of Com osition Column Wash (CH3OHzH20) on Recovery of 6 -Hydro>~. ,. dexamethasone ................. 38 B. Oxidation ........................................ 40 1. Choice of Oxidation Reagent .......................... 4O 2. Optimization of FCC Oxidation Conditions ................. 51 3. Determination of the Yield of 11,17-Keto Dexamethasone ........ 54 4. The Effect of Extracted Plasma Matrix on the Oxidation of Dexamethasone .................................. 55 5. Oxidation of 6B-Hydroxydexarnethasone ................... 57 C. Removal of Excess Reagent and Isolation of Oxidized Dexamethasone and 6B-Hydroxydexamethasone ................. 61 1. Silica vs. Solvent Extraction .......................... 61 2. Comparison of Sephadex LII-20 and Silica .................. 6 1 3. Optimization of silica solid phase extraction of 11,17-keto dexamethasone and 6,11,17—keto dexamethasone .............. 62 111. Conclusion ......................................... 66 References .............................................. 67 CHAPTER 3. COMPARISON OF GC-MS AND MS/MS TECHNIQUES IN II. THE ANALYSIS OF PLASMA FOR DEXAMETHASONE ...... 69 Introduction ......................................... 69 Sample Preparation and Analysis of Plasma for Dexamethasone Using SIM . . .70 A. Experimental ...................................... 7 O 1. Materials and Sample Preparations ....................... 7 0 2. Adjustment of Instrumental Parameters .................... 74 B. Results and Discussion ................................ 77 C. Conclusion ....................................... 9O MS/MS Analysis of Plasma for Dexamethasone ................... 91 vi A. Experimental ...................................... 9 2 B. Results and Discussion ................................ 93 C. Conclusion ...................................... 101 References ............................................. 102 CHAPTER 4. INVESTIGATION OF AN OXIDATION BASED-ASSAY FOR 6B-HYDROXYCORTISOL IN URINE; COMPARISON WITH A CONVENTIONAL GC-MS METHOD ................. 103 I. Introduction ........................................ 103 A. 6l3-Hydroxycortisol and Enzyme Induction ................... 103 B. Methods for Determination of Urinary 6B-Hydroxycortisol .......... 104 C. Proposed ECNI-MS Method for Determination of 6B-Hydroxycortisol . . .105 H. Experimental and Results ................................ 107 A. Preliminary Study .................................. 108 l. Extraction of Urine ............................... 108 2. Oxidation ..................................... 108 3. GC-ECNI/MS and GC-EI/MS Analysis of Oxidized Urine Extract . . .109 4. Investigation of Sample Introduction by Direct Inlet Probe for ECNI/MS ................................... 110 5. Conclusion .................................... 112 B. Determination of 6B-Hydroxycortisol in Guinea Pig Urine ........... 114 1. Sample Preparation ............................... 114 2. Instrumental Analysis ............................. 116 3. Results and Discussion ............................. 119 C. Determination of 6B—Hydroxycortisol in Human Urine by GC-EI/MS, GC—ECNI/MS, and DIP-ECNI/MS ........................ 121 1. Sample Preparations .............................. 121 2. Deuteration Experiments ............................ 122 3. Instrumental Analysis ............................. 124 4. Results and Discussion ............................. 125 III. Conclusions ........................................ 126 References ............................................. 127 CHAPTER 5. SUMMARY ................................... 129 References ............................................. 133 vii LIST OF FIGURES Figure 1.1 Numbering system for steroid carbon skeleton ................. 9 Figure 1.2 ECNI mass spectrum of PFB-oxime, TMS-ether of testosterone. The base peak at, m/z 181, represents the pentafluorobenzyl anion. This figure was taken from S. J. Gaskell, "Analysis of Steroids" in Methods of Biochemical Analysis, vol. 29, D. Glick, ed., Wiley Interscience, New York, pp. 385-434, 1983 .................. 21 Figure 1.3 Structures of steroids having high electron capture responses. I and II result from the oxidation of the naturally occuring c,ompounds cholesterol and 6B- -hydroxycortisol. III, IV, and V result from the oxidation of the corticosteroid drugs prednisolone, fludrocortisone, and dexamethasone, respectively .............. 24 Figure 2.1 Oxidative conversion of dexamethasone to 11,17—keto dexamethasone ................................... 33 Figure 2.2 Reconstructed selected ion current profiles from GC-ECNI/MS analysis of oxidized dexamethasone, 13C6-2H3- dexamethasone, and 6[3_-hydroxydexamethasone isolated from plasma by a) solvent extraction and b) C18 solid- -phase extraction. Arrow indicates retention time of oxidized dexamethasone ................... 35 Figure 2.3 Structures of some chromium (VI) oxidants; chromic acid (1), chromyl acetate (2), chromyl chloride (3), tert-butyl chromate (4), chromium trioxide-pyridine complex (5), dipyridine chromium (VI) oxide (6), pyridinium chlorochromate (7), pyridinium dichromate (8), and 2,2’-bipyridinium chlorochromate (9). Adaptedfromreference(15)..........................42 Figure 2.4 Mechanism for the oxidation of alcohols with chromium (VI) oxidants ....................................... 44 Figure 2.5 a) Mechanism for the oxidative cleavage of the CC bond of a 1,2- diol and b) proposed mechanism for the oxidative cleavage of the dexamethasone C-l7 sidechain ......................... 46 Figure 2.6 Oxidation of 6B-Hydroxydexamethasone to 6,11,17-keto dexamethasone ................................... 58 Figure 2.7 Mass spectra of oxidized 6B—hydroxydexamethasone obtained by a) electron impact ionization and b) electron capture negative ionization ...................................... 59 viii Figure 3.1 Selected ion current profiles for detection of the TMS-enol-TMS derivative of dexamethasone within plasma matrix, by GC—EI/MS .................................... 7 1 Figure 3.2 Selected ion current profile for detection of oxidized dexamethasone in plasma matrix, by GC-ECNI/MS ............. 72 Figure 3.3 ECNI mass spectrum of 11,17-keto dexamethasone and structure of 11,17-keto dexamethasone .......................... 78 Figure 3.4 GC-ECNI/MS selected ion current profile from 160 fg of 11,17- keto dexamethasone standard .......................... 79 Figure 3.5 Calibration curves for determination of dexamethasone in plasma b GC-ECNI/MS with SIM using a) 35 ng, and b) 3.5 ng of 1 C5-2H3-dexamethasone internal standard .................. 81 Figure 3.6 Selected ion current profiles from GC-ECNI/MS analysis of plasma sample P5 from patient undergoing the DST ............. 85 Figure 3.7 GC—ECNI/MS selected ion current profiles at m/z 310, 319, and 324, representing the (M-HF)- ions of oxidized dexamethasone, 13C6-2H3-dexamethasone, and 6B-hydroxydexamethasone, respectively, in plasma sample P5 ........................ 86 Figure 3.8 Daughter ion spectrum of the molecular anion of 11,17-keto dexamethasone obtained by ECNI/MS/MS with CAD ............ 94 Figure 3.9 The abundance of the (M-HF)- ion, m/z 310, from CAD of the molecular anion of 11,17-keto dexamethasone at different collision offset energies and collision gas pressures ................... 96 Figure 3.10 Comparison of selected ion current profiles obtained by a) GC- ECNI/MS using SIM, with those obtained by b) GC-ECNI/MS/MS with SRM, for the detemination of dexamethasone in a real plasma sample ........................................ 98 Figure 3.11 Calibration curves for determination of dexamethasone in plasma b GC-ECNI/MS/MS with SRM using a) 35 ng, and b) 3.5 ng of 1 C6-2H3-dexamethasone internal standard .................. 99 ix Figure 4.1 a) Oxidative conversion of 6B-hydroxycortisol to electrophilic product. b) ECNI mass spectrum of oxidized 6B- hydroxycortisol .................................. 106 Figure 4.2 Reconstructed total ion current chromatogram and mass chromatogram at m/z 314 from a) GC-EI/MS and b) GC- ECNI/MS analyses of an oxidized urine extract ............... 111 Figure 4.3 Selected ion current profile obtained by DIP-ECNI/MS analysis of an oxidized urine extract. The molecular anion of oxidized 6B- hydroxycortisol corresponds to m/z 314 ................... 113 Figure 4.4 Sample preparation scheme for the determination of 6B- hydroxycortisol in urine by MO-TMS derivatization and by oxidation ...................................... 115 Figure 4.5 Selected ion current profiles obtained from a) GC-EI/MS analysis of MO-TMS derivatized guinea pig urine sample and b) GC- ECNl/MS analysis of oxidized guinea pig urine sample .......... 117 Table 2.1 Table 2.2 Table 2.3 Table 2.4 Table 2.5 Table 2.6 Table 2.7 Table 2.8 Table 2.9 Table 2.10 Table 2.11 Table 2.12 Table 2.13 Table 2.14 Table 3.1 LIST OF TABLES Extraction recoveries of dexamethasone (dex) and 6B: hydroxydexamethasone (6B-OH dex) vs. percentage of methanol in solid phase extraction wash ......................... 3 9 Relative yields of 11,17-keto dexamethasone obtained with three different oxidation reagents ........................... 4 3 Reaction time vs. yield of 11,17-keto dexamethasone by PVPDC oxidation of dexamethasone at 70'C, in cyclohexane ............ 47 Relative yield of 11,17-keto dexamethasone by PVPDC oxidation vs. presence of excess H20 and heating and stirring. 26 ug of dexamethasone was used for each oxidation .................. 49 Relative yield of 11,17-keto dexamethasone by PDC oxidation vs. time ......................................... 50 The effect of a base on the yield of 11,17-keto dexamethasone ....... 52 The effect of heating on the yield of 11,17-keto dexamethasone ...... 53 Yield of 11,17-keto dexamethasone by PCC oxidation vs. time ...... 55 Ratio of GC-FID responses for 11,17-keto dexamethasone and the internal standard, 4-androstene-3,6,17-one, for oxidations conducted with and without presence of plasma matrix extract. 8: std. deviation of 3 injections ......................... 56 Relative yield of 6,11,17-keto dexamethasone vs. time ........... 60 Recovery of 11,17-keto dexamethasone by Sephadex LH-20 and by silica SPE column ............................... 62 HPLC analysis of fractions from silica column used for removal of FCC following oxidation of dexamethasone .................. 63 DIP-ECNI/MS analysis of silica SPE column for recovery of 11,17-keto dexamethasone and 6,11,17-keto dexamethasone ........ 64 GC-ECD analysis of silica fractions for isolation of 6,11,17-keto dexamethasone ................................... 65 Determination of dexamethasone in 1 ml blank plasma "spiked" with 2.09 ng of dexamethasone by GC-ECNI-MS with SIM ........ 84 xi Table 3.2 Table 3.3 Table 4.1 Table 4.2 Comprison of concentrations of dexamethasone in ng/ml of plasma as determined by GC-ECNI-MS with SIM and by conventional GC-EI/MS of the TMS-enol-TMS derivative ................. 87 Comparison of concentrations of dexamethasone obtained by GC- ECNI-SIM, GC-ECNI-SRM, and DIP-ECNI-SRM ............. 100 Comparison of GC-EI/MS of the MO-TMS derivatived urine with GC-ECNI/MS, and DIP—ECNI/MS, of the oxidized urine, for the determination of 6B-hydroxycortisol excretion by control and indu- *d guinea pigs ............................... 119 Comparison of GC-EI/MS of MO-TMS derivatives with GC- ECNl/MS and DIP-ECNI/MS of oxidized urine samples for the determination of the total daily excretion of 6B—hydroxycortisol by a 49-year old male ................................ 125 xii SIM SRM MS MS/MS CAD TQMS Q1 Q2 Q3 EF EI CI NCI EC ECNI DIP TIC APC S/N s/n s/b LOD GC KEY TO ABBREVIATIONS selected ion monitoring selected reaction monitoring mass spectrometry tandem mass spectrometry collisionally activated dissociation triplequadrupole mass spectometer first quadrupole second quadrupole third quadrupole high resolution electric field electron impact chemical ionization negative chemical ionization electron capture electron capture negative ionization direct inlet probe total ion current automatic pressure controller signal-to—noise signal-to-noise signal-to-background limit of detection gas chromatography xiii LC HPLC TLC ECD UV C1 8 SPE RIA PCC PDC PVPDC PyrzCr03 N aOAc BSTFA TMS MO-TMS TBDMS PFB DHEA LTB4 DST liquid chromatography high performance liquid chromatography thin layer chromatography flame ionization detector electron capture detector ultra-violet octadecylsilyl solid-phase extraction radioimmunoassay pyridinium chlorochromate pyridinium dichromate polyvinyl pyridinium dichromate pyridine chromium trioxide complex sodium acetate bis-(trimethylsilyl)-trifluoroacetamide trimethylsilyl methoxime trimethylsilyl ether tert-butyl-dimethylsilyl pentafluorobenzyl dehydroepiandrosterone leukotriene B4 dexamethasone suppression test xiv CHAPTER 1. INTRODUCTION 1. Mass Spectrometry in Biomedical Research It may be said that progress in science is limited by the development of analytical methodology by which the system of interest may be measured. For-instance, in the field of biochemistry, the chemical details of many biological processes have been discovered through the determination of the molecules involved in these processes. The application of mass spectrometry to this field has contributed greatly to the elucidation and quantitation of many compounds involved in biochemical systems. In order to use mass spectrometry for new applications, new methodology must often be developed. This thesis project involves the development of new methodology for the mass spectrometric determination of corticosteroids in biological samples. A. Advantages of Mass Spectrometry 1. Specificity Mass spectrometry is a very sensitive and an inherently specific technique. The specificity arises from the fact that the masses of ions produced from a molecule are measured directly. The basis for many other analytical techniques is the measurement of a physical pr0perty, such as the absorption of radiation. The structural moiety responsible for an absorption of radiation may be common to many different molecules, whereas the masses of the ions produced from a compound, and the relative abundances of these ions, are very specific for the compound. In mass spectrometry, a molecule is introduced into the vapor phase and bombarded with a relatively high energy (70 eV) beam of electrons. The collision l 2 between the molecule, M, and the electron imparts so much internal energy to M that in order to dissipate this energy, the molecule first ejects an electron, resulting in the formation of a molecular ion, M+. Molecular weight information is obtained by observing the mass-to-charge ratio at which the molecular ion appears. The molecular ion often fragments further to relieve itself of the remaining excess internal energy. The fragment ions produced are indicative of the structural features of the molecule. A mass spectrum is simply a plot of the mass-to-charge (m/z) ratio of the ions against their relative intensities. Except in the case of some isomeric compounds which produce similar to identical mass spectra, a mass spectrum is a highly specific ’fingerprint’ for a compound, providing a means by which compounds can be positively identified. There are several types of mass spectrometers commercially available. Currently the most popular types of mass spectrometers are the quadrupoles and the double focusing mass spectrometers. Quadrupole mass spectrometers consist of an ion source, focusing lenses, a quadrupole mass analyzer, and a detector. The mass analyzer consists of four cylindrical rods. DC potentials of opposite sign are applied to the two sets of adjacent rods. An RF potential is superimposed on the DC potentials. The ratio of the DC and RF potentials determines the mass resolution of the quadrupole mass analyzer. The magnitudes of the DC and RF potentials can be adjusted such that only ions of a specified m/z will have a stable trajectory through the quadrupole. The other ions are filtered out. The DC and RF potentials can be varied so that ions of each m/z value are sequentially allowed to pass through the quadrupole and be detected, resulting in a mass spectrum. A variation on quadrupole instruments is the triple quadrupole mass spectrometer (TQMS). The TQMS is identical to a single quadrupole mass spectrometer except that its mass analyzer consists of three quadrupole mass filters in tandem. With this type of instrumental arrangement, two stages of mass analysis can be performed; hence, the TQMS performs a tandem mass spectrometry experiment, often called by the acronym 3 MS/MS. In MS/MS, the first quadrupole can filter out all the ions produced from a sample except the selected parent ion. The parent ion is passed into the second quadrupole, which serves as a high transmission collision cell, where the ion may undergo collisionally activated dissociation (CAD). The ions produced upon CAD are called daughter ions. All the daughter ions are transmitted into the third quadrupole, which is then scanned to obtain the mass analysis of the daughter ions. Double-focusing mass spectrometers consist of an electrostatic analyzer and a magnetic analyzer. Ions are accelerated to high energy (3-10 kV) and are directed into the electrostatic analyzer which acts as an energy filter by allowing all ions of a given kinetic energy to pass into the magnetic sector which then transmits the ions to the detector based on their momenta. The resolution of double-focusing mass spectrometers may be controlled by adjusting the width of the slits through which the ions must pass. As the slits at the exit of the ion source and the exit of the electric sector are narrowed, the energy spread of the ions which are allowed to pass is decreased, resulting in much higher resolution. The type of mass spectrometer chosen for a particular analysis depends on the required specificity. TQMS adds an extra dimension of selectivity to mass spectrometric analysis by differentiating ions of a given nominal mass on the basis of their daughter ion spectra. Double-focusing mass spectrometers operated at high resolution are capable of differentiating ions of the same nominal mass by more accurately measuring the exact rn/z values of the ions. Both of these types of mass spectrometric methods have been used to provide more specificity for the analysis of analytes in complex biological matrices 2. Sensitivity Mass spectrometry is useful for biomedical research not only due to its high specificity, but also to its high sensitivity which allows trace determinations to be performed. Often, the analyte to be measured in a biological sample may have a very low or ’trace’ (ng/g) concentration. A mass spectrum having good signal-to—background ratios may be obtained from as little as 1 ng of a pure material. In the presence of high background (chemical "noise") from the sample matrix, more of the analyte must be present to obtain a mass spectrum having sufficient signal-to-background ratio to provide interpretable information for structural elucidation. If greater sensitivity is required for a quantitative mass spectrometric assay, the technique of selected ion monitoring (SIM) may be used. In this technique only the ion ’ currents from selected mass-to-charge ratios that are indicative of the analyte of interest are monitored. Since a longer time is spent monitoring these ion currents, the signal-to- background (s/b) ratios for these ions are greater than they would be if the whole mass range were scanned. Because the s/b ratios are greater, the detection limit is lowered. Typical detection limits of EI-MS assays using SIM are in the low ng/ml range (1-100 pg absolute quantities introduced) ( l ). B. Some Applications of Mass Spectrometry 1. Qualitative Gas Chromatography-Mass Spectrometry The use of mass spectrometry has become increasingly important in biomedical research over the last two to three decades, especially since the development of gas chromatography-mass spectrometry (GC-MS) in the 1960’s. The coupling of the gas chromatograph with the mass spectrometer allowed the identification of the components 5 of complex mixtures. Even after several isolation procedures, samples of biological origin are likely to contain many compounds other than the compound of interest to the investigator. Gas chromatography separates mixture components based on their differential boiling points and affinity for the stationary phase of the chromatographic column. Ideally, the components enter the mass spectrometer sequentially, separate from all interfering substances which were isolated along with the component(s) of interest. In this way, a mass spectrum representing the pure component may be gathered. Mass spectrometry has been used for the identification of precursors and metabolites of biochemical pathways. For instance, leukotrienes (2) and their metabolites (3) have been both confirmed and discovered through the use of mass spectromety. Other examples of biologically important molecules identified by mass spectrometry include metabolites of biogenic amines (4), peptides (5), and xenobiotically modified analogs of nucleic acids (6). Mass spectrometry is also used extensively in drug metabolism studies. The main objectives of drug metabolism studies are to identify the pathways by which drugs are modified in the body and to determine the quantitative importance of each intermediate and pathway (7). GC-MS is particularly suited for drug metabolism studies because it allows reliable determination of many similar compounds within a single analysis. Drug metabolism studies by less specific methods are often unsuccessful because the similarity of a drug and its metabolites makes differentiation of the compounds difficult. 2. Quantitative Mass Spectrometry Used as a quantitative technique, mass spectrometry has, in addition to high sensitivity, another important advantage over many other quantitative methods in that isotopically-labelled internal standards can be used. The main advantages of using isotopically-labelled internal standards is the improved accuracy and precision of the 6 assay because the behavior of the internal standard is identical to that of the analyte throughout isolation and analysis procedures (8). Some examples of the uses of quantitative mass spectrometry are in the determination of the pharmacokinetics and bioavailability of drugs. Bioavailability, defined as the amount of administered drug that becomes available for pharmacological effect, differs considerably depending on the formulation of the drug. Bioavailability is best determined by measuring the plasma levels of the drug. The complexity of the sample and the low concentration of the drug warrant the use of mass spectrometry (9). Drug pharmacokinetics, the rate of metabolism and/or disposition of a drug, also can be studied using mass spectrometry (7). The circulating concentrations of drugs and/or their metabolites are monitored over time. Such information is important for determining proper dosage. Also, drugs may be administered and monitored to test and study a biological function, for instance, enzyme induction or endocrine function. 3. Mass Spectrometry as a Reference Method Another function of mass spectometry in biomedical research is to serve as a standard against which the reliability and accuracy of other new methods, such as immunoassays, may be tested (10). Immunoassays are based on the specific binding of an antigen, which is the analyte of interest, to an antibody. Antibodies are proteins that are formed as an immunologic response to a foreign substance. To obtain the antibodies for an immunoassay, either an animal or a cell culture is exposed to the substance for which the immunoassay is to be developed. Antibodies are produced by the animal or cell culture as a response to this substance. If a cell culture is used to produce the antibodies, they are termed monoclonal antibodies because they result from a single kind of cell. After isolation and purification of the antibodies, an assay for the antigen may be developed based on competitive binding of the antigen and a labelled analog of the 7 antigen. For instance, in radioimmunoassay a known amount of radiolabelled antigen is incubated with an amount of antibody. This preparation is then re-incubated with the sample which contains the unlabelled antigen to be assayed. A portion of radiolabelled antigen will be displaced by the unlabelled antigen present in the sample. The antigen- antibody complex is then separated from the non-bound matrix. The amount of radiolabelled antigen displaced into the non-bound matrix is proportional to the amount of unlabelled antigen that was present in the sample. The reliability of RIA and other types of imrnunoassays is dependent upon the specificity of the antibody. Interaction of the antibody with components of the sample matrix that are similar to the analyte, such as its metabolites or isomeric compounds or analogs, is called cross-reactivity. Cross-reactivity obviously leads to false results. However, immunoassays have the advantages of very high sensitivity, simple or no sample preparation. Also, analyses can be done on very small amounts of sample, and once a specific antibody is raised, multiple samples can be analyzed very rapidly using devices such as scintillation spectrometers. Because of these advantages, it is desirable to be able to use these methods provided they can be validated. Mass spectrometry is often used to validate immunoassay methods. Some recent examples include comparisons of SIM and RIA for the measurement of l7l3-estradiol (11) and the quantitation of 5-hydroxy eicosatetraenoic acid (12). 11. Conventional Mass Spectrometry for Corticosteroids Analysis Mass spectrometry has been used for the structural elucidation and quantitative determination of many classes of biological molecules. This dissertation is concerned with the deve10pment of methodology for the determination of corticosteroids by mass spectrometry. This section contains background information on the corticosteroids and the conventional methods available for their determination. A. Corticosteroids Corticosteroids are a class of steroid hormones secreted by the adrenal gland. These steroids have 21 carbon atoms and a 4-ene-3-ketone moiety. The basic structure of a steroid is shown in Figure 1.1. The types of steroids differ by the number of carbons within the basic steroid skeleton. Nomenclature rules are also summarized in Figure 1.1. The corticosteroids are responsible for regulating many biological phenomena including carbohydrate, protein, and lipid metabolism, and water and electrolyte balance. Corticosteroids also influence the functions of the kidney, skeletal muscle, the nervous system, and the cardiovascular system. The corticosteroids can be further categorized as mineralocorticoids, which are responsible primarily for maintaining electrolyte balance, and glucocorticoids, which are primarily involved in carbohydrate metabolism, although there is some overlap of activity between these two categories. The glucocorticoids are also potent anti-inflammatory agents. In addition to cortisol (4-pregnen-11j3, 17a ,21- triol-3, 20-dione), which is the most potent naturally occurring glucocorticoid, synthetic analogs of cortisol having greater potency have been developed for clinical use. For instance, dexamethasone (90t-F, 16a—CH3, 1-ene analog of cortisol) has 25 times the anti- inflammatory effect of cortisol and no effect on electrolyte metabolism. The isolation, structural elucidation, and economical synthesis of corticosteroids has made the clinical use of these compounds possible. Identifications of the structures of corticosteroids were postulated based on observation of physical properties such as ultra-violet absorption, molecular rotation, and infra-red absorption, and chemical reactivity toward oxidation and reduction or organic reagents selective for specific types of functional groups. Confirmations of the postulated structures were done by comparing the properties of the unknown with the properties of a synthesized version of the postulated structure. Confirmation of postulated structures was made much simpler with the development of GC—MS. 21 22 24 26 18 20 3 25 12 17 11 13 16 19 1 14 ’15 2 8 3 7 5 4 6 A = androstane, C19 skeleton P = pregnane, C21 skeleton C = cholestane, C27 skeleton ene = double bond 01 = hydroxyl group one = carbonyl group or = substituents extending into the plane of the page B = substituents extending out of the plane of the page Figure 1.1. Numbering system for steroid carbon skeleton. 10 After the structures of the corticosteroids had been determined, there developed more concern for quantitiative measurement of corticosteroids. The quantitative determination of the naturally occurring corticosteroids is done primarily to assess the function of the adrenal cortex. Quantification of the synthetic analogs used as drugs is done primarily to assess proper therapeutic levels. A variety of methods are available for the quantification of both the endogenous and synthetic corticosteroids. These include colorimetric, fluorimetric, and gas and liquid chomatographic, as well as mass specuometric methods. Colorimetric and fluorimetric methods have been in use the longest, but due to the prevalence of non-specific interferences in these types of methods and the development of more specific chromatographic methods, more laboratories are using GC and LC based methods. Chromatographic based methods have greater specificity than the colorimetric and fluorimetric procedures because identification is based not only on detector response of separated components, but also on retention times of the separated components. Chromatographic methods also have the advantage that identification and quantification of multiple sample components may be accomplished within a single analysis. Although preliminary identification of chromatographic peaks can be made based on retention time data, confimation of peak identity must often be done with GC-MS, especially when the component(s) in question is minor or incompletely resolved from other components. It is possible to make positive identifications as well as quantitative detemrinations of both co—eluting components and minor components using GC-MS due to the specificity of the mass spectrometer. B. Derivatization of Corticosteroids for Analysis by GC-MS In order to determine corticosteroids by GC-MS, the polar groups on the molecule must be modified to render these groups non-polar, allowing the compound to vaporize at a temperature lower than that at which the compound would decompose. The presence 1 l of a Nor-hydroxy on the C-17 sidechain of corticosteroids results in thermal lability (13). Derivatization stabilizes reactive structural arrangements and prevents thermal decompostion. The vapor phase properties of cortisteroids are also improved by derivatization of the polar groups, resulting in good gas chromatographic behavior. Criteria for a good derivative include the following (14): l) completeness of derivative formation 2) selectivity of derivative formation 3) ease and reliability of derivative formation 4) stability of the derivative that is formed 5) GC retention time and chomatographic quality (symmetry) of the peaks 6) the stability and inertness of the derivative during GC Also, when SIM is used, it is important to employ a derivative that provides high abundance of the molecular ion or other structurally-informative ions in the mass spectrum. Different derivatives have been developed to try to meet these criteria; however, the most commonly used derivative for general steroid analysis is the methoxime (MO) - trimethylsilyl (TMS) ether (MO-TMS). Reaction with methoxyamine hydrochloride converts ketones to methoxirnes. In this protected form, the ketones will not enolize during the subsequent reaction with a reagent such as bis(trimethylsilyl)- trifluoroacetamide (BSTFA) which converts the hydroxy groups to TMS ethers. MO- TMS derivatives have a number of advantages and disadvantages. The mass spectra of MO-TMS derivatives of steroids are characterized by losses of 90 amu (TMSOH) and 31 amu (OCH3 radical). These characteristic losses aid in interpreting the spectrum making it relatively simple. The abundance of the molecular ion and the high mass ions is relatively low, however, decreasing the sensitivity available for SIM assays. Because l 2 MO-TMS formation is a two step process, a mixture of products can occur. Also, because of the steric hindrance of the 17a-hydroxy of corticosteroids, longer reaction conditions are required to fully derivatize these steroids. The corticosteroid drug, dexamethasone, possesses a 16a-methyl group which makes MO-TMS formation particularly difficult. Three hours are required for methoxime formation and six hours for TMS ether formation at elevated temperatures (15). The trimethylsilyl ether (TMS)-enol-TMS ether (16) has been found to be most useful for the analysis of adrenocortical steroids. Enolization of the C—20 ketone followed by TMS ether formation allows a one step reaction for the analysis of corticosteroids having different sidechain types. This reaction takes less time than MO- TMS formation. Some other derivatives available for corticosteroids include various types of silyl ethers such as the t-butyl dimethyl silyl (T BDMS) ether, and various cylic derivatives such as the methyl boronate (17), dimethylsiliconide (18), and the diethylhydrogensilyl cyclic diethylsilylene (19) derivatives, which are specific for 1,2- or 1,3-diols. The TBDMS derivative has an abundant high mass fragment in its mass spectrum which is advantageous for SIM assays. However, this type of derivative is not really practical for corticosteroids because the molecular weight of the derivatized compound will be very high if the multiple hydroxy groups are derivatized. Also, MO formation is still required and the TBDMS group is even more bulky than a TMS group making derivatization of sterically hindered groups even more difficult. The methyl boronate derivatives afford abundant molecular ions of relatively low mass. Also the GC retention times of these corticosteroid specific derivatives are relatively short compared to MO-TMS derivatives. However, subsequent trimethylsilylation of unprotected hydroxy groups can lead to partial cleavage of the boronate (20). One disadvantage common to all these derivatives is that they are all subject to hydrolysis. Therefore, storage conditions must be moisture free. 13 C. Other Mass Spectrometric Techniques for Corticosteroids Analysis The determination of very low levels of specific corticosteroids within a complex biological matrix sometimes requires even higher specificity and sensitivity than can be obtained through the use of GC-MS with SIM. Some instrumental methods have been used to increase the sensitivity and/or selectivity for corticosteroid determination. These include high resolution selected ion monitoring (HR-SIM) and alternate ionization techniques such as positive chemical ionization (CI), negative chemical ionization (NCI), and electron-capture negative ionization (ECNI). High resolution mass spectrometry (HRMS) is not practical for many investigators due to the expense of a high resolution mass spectrometer. However, there are some groups which have made use of this technique and achieved high sensitivity due to the increased specificity of the technique. For example, the selectivity for 1713- estradiol in plasma is shown to be considerably increased through the use of HR-SIM (21). Also the improved selectivity of HR-SIM allowed 2 pg/ml level detection of testosterone in the saliva of female subjects (22). Positive CI may offer a sensitivity advantage over electron-impact ionization in selected ion monitoring assays since the relative intensity of the molecular species to be monitored is increased. Upon chemical ionization, little energy is imparted to the molecule, therefore little or no fragmentation of the molecular species (M+H)+ occurs. For certain molecules, NCI and ECNI offer both sensitivity and selectivity advantages. This dissertation emphasizes ECNI-MS as a technique for enhancing the selectivity and sensitivity of assays for corticosteroids. 14 III. Electron Capture Negative Ion Mass Spectrometry A. History and Theory of Electron-Capture Negative Ionization Much of the early investigation of electron-capture negative ionization (ECNI) of organic compounds was done in the laboratory of von Ardenne (23, 24). His experiments were conducted with an argon discharge source in which a primary electron beam was slowed by collisions with a high pressure of argon atoms resulting in the formation of a low energy plasma of electrons and argon ions. Molecules having positive electron affinities are able to absorb a low energy or ’thermal’ electron resulting in the formation of a molecular anion. After absorption of the electron, the molecular anion is in an excited state, having internal energy at least equal to its electron affinity. The excess 9 internal energy of the excited molecular anion may be dissipated through collisions with inert species or by radiative emission. If the molecule anion has a sufficient number of internal degrees of freedom over which to dissipate the excess energy, the anion will be sufficiently long-lived for stabilization to occur. In small molecules with few internal degrees of freedom, ejection of the electron is more likely to occur before stabilization can occur. Molecular anions are formed by electron capture of electrons having energies in the 0 to 2 eV range. If the energy of the captured electron is larger, dissociative electron capture can occur. In a collision with a higher energy electron (0-15 eV) more internal energy must be dissipated in order to stabilize the anion. The internal energy absorbed upon electron capture is dissipated by breaking one or more bonds, resulting in a stabilized fragment anion. The "driving force" for an electron capture reaction may be the formation a resonance stabilized fragment ion, such as (M-H)- (52). The argon discharge source used by von Ardenne was an in-house modification. When it was found that the high pressure source used for positive CI could also be used 1 5 for ECNI more laboratories began to investigate this mode of ionization. The mechanism of the formation of the low energy electron plasma under CI conditions is shown by the equation CH4 + e' (100 eV) --> CH4+ + e-(t) + e' (<100 eV) where e-(t) represents an electron having only thermal energy. Upon each ionizing collsion, the primary electron loses 30 eV (25). After several collsions, the primary electrons are thermalized. In the high pressure CI source, a plasma of thermalized electrons is readily produced. Dougherty (26) was the first to suggest the use of the conventional CI source for obtaining electron-capture negative ionization spectra. He used this technique to obtain ECNI spectra of chlorinated environmental contaminants (49). Hunt et al (27) were among the first groups to anticipate the analytical potential of ECNI for biological applications. A detection limit of 25 fg (s/n = 4) of a pure standard of the pentafluorobenzoyl derivative of amphetamine was obtained by ECNI-MS with SIM. Since then, applications of ECNI have increased. It is commonly used for detection of many halogenated environmental contaminants as well as for the detection of compounds of biological interest such as prostaglandins and leukotrienes derivatized as pentafluorobenzyl esters (28). Methods have been developed for many other compounds such as neurotransmitters (29) and drugs (30). B. Limitations of Electron Capture Negative Ionization Limitations of ECNI include the acute dependence of the extent of ionization on the ionization parameters, such as temperature and pressure, as well as the limited applicability of the technique. Because the mean energy of the electrons in the CI source is highly dependent on the temperature and pressure of the CI gas, and the extent of 1 6 ionization and fragmentation of an analyte is dependent upon the energy of the electrons, it is important to carefully regulate both temperature and pressure. This is particularly important when conducting quantitative analyses. Slight pressure changes from run to run decrease precision. Even pressure changes during a single analysis can decrease precision if a structural analog is used as internal standard. Better precision is maintained when isotopically labelled internal standards are used since any flucuations in conditions are experienced simultaneously by the analyte and internal standard as they co-elute. The purity of modifying gas is also of utmost importance for ECNI. The presence of trace amounts of 02 or H20 within the mass spectrometer can result in ion/molecule reactions with 02, 02', 0-, or OH- which may be more facile than the electron/molecule reactions (31). For example, trace amounts of oxygen within the modifying gas have been demonstrated to completely change the ECNI spectrum of the plant hormone methyl abscissate (32). Normally, the CH4-ECNI mass spectrum of methyl abscissate contains m/z 278 (M') as the base peak and a few minor fragment ions. If a trace of oxygen is introduced, an ion/molecule reaction occurs to form an (M+02)- adduct ion which then undergoes unique fragmentation pathways. In this case, the spectrum consists of the ions resulting from electron capture as well as the ions resulting from the ion/molecule reaction, but the peaks dominating the spectrum are fragments resulting from the ion/molecule reaction. The cleanliness of the ionization chamber may also affect ECNI spectra. Halogen impurities on the chamber walls can cause changes in spectra such as the appearance of M+Cl adducts (33). Such surface reactions with species adsorbed to the source walls can also distort the chromatographic peak shape producing broadening and tailing of peaks (34, 35). The care that must be taken to control ion source ionization parameters makes the use of ECNI more difficult than EI ionization, which has become rather routine. Still, the reproducibility of ECNI can be reasonably good as long as care is taken concerning l 7 the factors mentioned above. Stemmler and Hites (36) published a study of the parameters which affect the reproducibility of ECNI spectra which indicated thatthese spectra are more reproducible than has been commonly perceived. They found that with careful control of instrumental parameters, such as ion source pressure, emission current, ion source temperature, and ion focus potential, reproducible ECNI spectra could be obtained throughout a one-year period. An inter-laboratory study by the same authors supported these conclusions (37). Although exact duplication of the relative intensities of the ions in the ECNI spectra would require assessment and control of the lens and ion source parameters, and of the presence of impurities, the general features of the ECNI spectra of several compounds were reproduced on different instruments. C. Advantages of ECNI The two notable advantages of ECNI are the increase in sensitivity and the selectivity of the ionization. 1. Sensitivity For certain types of compounds, electron capture negative ionization can be orders of magnitude more efficient than positive ionization. For instance, the total ion current obtained from 2 n g of the PFB-TBDMS derivative of LTB4 is 50 times greater by ECNI than by El ionization and 20 times greater than by positive CI (38). Hunt used pulsed positive negative ion chemical ionization mass spectrometry to demonstrate that negative ion currents were 100 to 1000 times greater than the corresponding positive ion currents for compounds such as anthraquinone, phthalic anhydride, and trinitrotoluene (23). The greater negative ion currents can be explained on the basis of the rate constants for positive ion formation and negative ion formation by electron capture. As an l 8 approximation, the ratio of negative ion current (d(M’)/dt) to positive ion current is proportional to the ratio of the rate constant for an electron/molecule reaction to the rate constant of a positive reagent ion/molecule reaction (50). The maximum values for the rate constant of reactions between reagent ions and molecules are on the order of 3-4 x 10-9 cm3 mol-l sec-1. In contrast, electron/molecule reaction rate constants are on the order of 4 x 10'7 cm3 mol-l sec-1 (51). The higher mobility of the electrons in the chemical ionization plasma may account for the difference in rate constants. Based on these values, negative ion current could be two orders of magnitude greater than positive ion current. Even if the actual ionization of the analyte is not enhanced over its positive ionization efficiency, sensitivity of an assay may still be enhanced by ECNI. The effective signal-to-background ratio is normally enhanced for a compound with good electron capture response because the ionization of most sample matrix components will have poor or no electron capture response. Also, because ECNI is a ’soft’ ionization technique, little fragmentation occurs, allowing the ion current to be concentrated in a few ionic species. This effect maximizes the sensitivity available with SIM. 2. Selectivity ECNI is called a selective mode of ionization because most compounds do not give an appreciable response under ECNI conditions. Among the few classes of compounds which have good EC properties are those that have a halogen or a nitro group attached to a pi-bonded carbon system, a highly conjugated carbon system, or a weak electron capturing group, such as a carbonyl, attached to a conjugated system (30). For compounds falling into these classes, the fact that few other compounds respond under these conditions can be used to advantage for the determination of these compounds in complex sample matrices. For example, Dougherty shows that dieldrin at 0.1 ppm 1 9 concentration in chicken tissue extract is easily identified by ECNI. The most prominent peak in the spectrum represents the molecular ion of the chlorinated pesticide dieldrin. However, under positive ionization conditions interferences from co-extracted lipids dominate the spectrum obscuring identification of the dieldrin (39). Other examples of compounds which have been detected in complex matrices by ECNI are the benzodiazepines. This class of drugs contains a halogen atom on a conjugated ring system and, thus, has good electron capture response. The ECNI assay developed for fluazepam, a particularly potent benzodiazepine hypnotic drug, was the first chemically based method that was sensitive enough to detect fluazepam to provide a pharmacokinetic profile in man (40). D. Derivatization to Enhance ECNI Response Because determination of compounds of biological interest are normally found in complex matrices, it would be an advantage to be able to selectively ionize the ' component(s) of interest without ionizing the considerable matrix components. Because most compounds of biological interest are neither highly conjugated nor halogenated, the approach of derivatization of functional groups with fluorinated moieties has been applied to enhance sensitivity and selectivity of detection of analytes within biological matrices. Although the electron capture responses of the halogens increase from fluorine to iodine, the use of fluorine is the best compromise for a derivative because the increase in retention time is least when fluorines are used (41). Some examples of derivatives include pentafluorobenzyl esters, perfluoroacyls, bis(trifluoromethy1) benzoyls, and pentafluorobenzyloximes. Assays for fatty acids, amino acids, steroids, and other types of molecules based on derivatization with the appropriate reagent have been published. Recently, a 2 fg limit of detection (LOD) was reported for PFB ester-dimethylethylsilyl ether derivatives of bile acids (42). 2 0 Despite the success with PFB esters, limits of detection for many ECNI assays based on derivatization represent only modest gains over those of the corresponding EI- MS based methods. The expected gains in sensitivity and selectivity may not be realized in many cases without extensive analyte purification because of two reasons. The first is that these types of derivatization reactions are not very selective. Any concomitant in the sample matrix having a reactive functional group will produce a highly responding derivative. For instance, the perfluoroacyl derivatives can be formed with a variety of functional groups including hydroxyls, amines, phenols, and the enol form of ketones. Thus, the effective s/b is not increased without extensive purification of the analyte. The second reason for lower than expected sensitivity in the use of derivatives is that many derivatives used to enhance electron capture result in the production of a mass spectrum that is more representative of the derivatization moiety than of the intact derivatized molecule. An example of this type of mass spectrum is shown in Figure 1.2, which was obtained from the PFB oxime-TMS ether of testosterone. Low abundance of the molecular ion or high mass fragments will decrease the sensitivity that can be realized in a quantitative assay based on SIM of these ions. The most intense ions which are unrepresentative of the intact derivatized analyte would not be suitable for a SIM assay because every derivatized molecule in the matrix will produce these same ions. 100* 50‘ relative abundance 100 Figure 1.2. 21 181 OSHCHag C5F5CH2EO-N m/z 181 <3 167 197 jl .l 1 11. .1 1‘ ' L LJ-l . 1 200 300 1.00 500 600 m/z ECNI mass spectrum of PFB-oxime, TMS-ether of testosterone. The base peak at, m/z 181, represents the pentafluorobenzyl anion. This figure was taken from S. J. Gaskell, "Analysis of Steroids" in Methods of Biochemical Analysis, vol. 29, D. Glick, ed., Wiley Interscience, New York, pp. 385-434, 1983. 22 Corticosteroid determinations by ECNI have been very limited because the derivatization of this class of steroids with electrophilic moieties has been unsatisfactory due to the multiple hydroxy and ketone groups present in'these types of molecules. For example, the formation of perfluoroacyl derivatives of corticosteroids, such as the heptafluorobutyryl, results in complex mixtures of products due to enolization of ketone groups, making these types of derivatives unsuitable for quantitative analyses. Another group of electron capturing derivatives for corticosteroids, the electron capturing boronates, such as the 3,5-bis(trifluoromethyl)benzene boronate, were investigated by Poole et a1 (20). Multiple products were obtained upon derivatization of cortisol and other related corticosteroids with 3,5-bis(trifluoromethyl)benzene boronic acid. Also, subsequent trimethylsilylation of the remaining hydroxy groups resulted in partial cleavage of the boronate. Therefore, these derivatives were not suitable for quantitative determination of corticosteroids. An ongoing goal of some researchers in analytical methodology is to develop new derivatives which will meet the needs for specific types of determinations, such as derviatives which may produce stable molecular anions and/or high mass fragments under ECNI conditions. A goal in this laboratory is to deve10p alternative approaches to analyte modification which will be more selective based on chemistry, and which will allow the advantages of sensitivity and selectivity of ECNI to be realized without the requirement of extensive analyte purification. IV. An Alternative Approach for Enhancing ECNI Response of Corticosteroids Of the derivatives commonly used to increase electron capture response, none seem to be suitable for the derivatization of corticosteroids. Therefore, an alternative approach, involving the chemical oxidation of corticosteroids, is being studied in this laboratory as a means for enhancing the electron capture response of these compounds. 23 A. Previous work It was first noted by Lovelock (43) that steroids containing cab-unsaturated ketones gave unexpectedly high ECD responses. The magnitude of the response of the steroid containing the 0t,B-unsaturated ketone increased depending on the presence and position of other ketones or double bonds within the molecule. Poole (20) confirmed that certain moieites, in specific geometries (especically the C-11 and C-17 ketones in combination with the 4-ene-3-ketone) on the steroid nucleus results in high EC response. Some of these structures are shown in Figure 1.3. The natural presence of electron capture properties in certain steroids has been used to advantage in some cases for the sensitive detection of these compounds in complex samples. For instance, an assay was developed to detect low levels of the steroid drug melengestrol acetate in meat (44). Also, ecdysteroids, a class of insect hormones that contains a 7-ene-6-one moiety, have been detected in insects at low levels using the electron capture detector (45). The use of chemical oxidation to incorporate electron capturing moieties into a steroid was found to be useful for the detection of dehydroepiandrosterone (DHEA, 5— androsten-3B-ol-l7-one) by de Jong and van der Molen (46). It was found that the CrO3 oxidation product of DHEA, in addition to having good gas chromatographic properties, was also highly electrophilic due to the 4-ene-3,6-dione moiety incorporated upon oxidation. This group developed an assay for DHEA in plasma based on oxidation of the extract, followed by GC-ECD. Specificity for DHEA was ensured by separation of DHEA from other 5-ene-3-ols prior to oxidation. The electron capture detector and ECNI operate on the same principle. In the ECD, a radioactive (63Ni or 3H) foil emits [3 particles which ionize the carrier and make-up gas (nitrogen or argon) resulting in a plasma of thermal electrons that are available for electron capture by compounds as they elute from the GC and pass through the detector along with the make-up gas. Most compounds having ECD response will also give a response to ECNI-MS. 24 Figure 1.3. Structures of steroids having high electron capture responses; I and 11 result from the oxidation of the naturally occuring compounds, cholesterol and 6B-hydroxycortisol. III, IV, and V result from the oxidation of the corticosteroid drugs prednisolone, fludrocortisone, and dexamethasone, respectively. 25 Previous work in this laboratory has shown the feasibility of using oxidation to incorporate C-11 and C-17 ketone moieties and enhance the EC properities of corticosteroids drugs (47). Oxidation of dexamethasone, fludrocortisone, fluormetholone, prednisolone, and flumethasone with chromium trioxide di-pyridine complex results in 11,17-keto analogs of these drugs that all have very high ECNI-MS responses. The feasibility of this approach as well as the advantages of sensitivity and selectivity were demonstrated by an assay for dexamethasone in horse urine (48). Ng/ml quantities were detected by ECNI-MS following oxidation of the solvent extract of the urine. Dexamethasone was oxidized to a highly responding structure, whereas the other co-extracted steroids were not. Most endogenous steroids, lacking any double bonds, are not oxidized to highly responding products. Only the corticosteroids, progesterones and some androgens have the required (LB-unsaturated ketone. And only a small percentage of the androgens and progesterones have an oxygen at the C-11 position. Fewer still have a C-l couble bond or are oxygenated at the 06 position. Thus oxidation provides a way to selectively enhance the responses of only a few types of steroids, namely those having the 4-ene-3-one-6-oxy moiety, the 1,4-diene-3-one-11-oxy moiety, or the 5-ene- 3-ol moiety. 26 B. Majors goals of this research The three major goals of the research project described in this dissertation are: l) the optimization of the oxidation methodology for the quantitative conversion of corticosteroids to their electrophilic oxidized analogs; 2) to ascertain the extent to which the enhanced selectivity of the oxidation and the ionization will allow a less selective method of sample introduction, such as direct probe rather than GC; and 3) to apply this oxidation methodology a) to the determination of dexamethasone and its major metabolite, 6B-hydroxydexamethasone, in plasma from patients undergoing the dexamethasone suppression test, and b) to the determination of 6B—hydroxycortisol in urine for use as an index of enzyme induction. 10. ll. 12. 13. 14. 15. 16. 17. 18. REFERENCES A. L. Burlingame, T. A. Baille, and P. J. Derrick, Anal. Chem, 58, 165R-211R, 1986. R. C. Murphy, S. Hammerstrom, and B. Samuelsson, Proc. Natl. Acad. Sci. USA, 76, 4275, 1979. D. O. Stene and R. C. Murphy, J. Biol. Chem, 263, 2773-2778, 1988. S. Higa and S. P. Markey, Anal. Biochem, 149, 492-500, 1985. F. De Angelis, M. Botta, S. Ceccarelli, and R. Nicoletti, Biochem. J ., 236, 609- 612, 1986. R. K. Mitchum, J. P. Freeman, f. A. Belank, and F. F. Kadlubar, in Mass Spectrometry in the Health and Life Sciences, A. L. Burlin game and N. Castagnoli, Jr., eds., Elsevier, Amsterdam, pp. 547-580, 1985. A. Goldstein, L. Aronow, and S. M. Kalman, Principles of Drug Action: The Basis of Pharmacology, John Wiley and Sons, New York, 2nd edition, p. 230, 1974. J. Sjovall and M. Axelson, Vit. and Hormones, 39, 31-144, 1982. W. A. Garland and M. L. Powell, J. Chromatogr. Sci, 19, 392-434, 1981. H. Aldercreutz, "Biomedical Application of the Mass Spectrometry of Steroid Hormones" in A. Quayle, ed., Advances in Mass Spectrometry, vol . 88, London, Heyden, 1165-1179, 1980. V. Papadopoulos, S. Carreau, E. Szerman-Joly, M. A. Drosdowsky, L. Dehennin, and R. Scholler, J. Steroid Biochem, 24, 1211-1216, 1986. J. Y. Wescott, S. Chang, M. Balazy, D. O. Stene, P. Pradelles, J. Maclouf, N. F. Voeikei, and R. C. Murphy, Prostaglandins, 32, 857-873, 1986. H. Gottfried, Steroids, 5, 385-397, 1965. T. A. Baillie, C. J. W. Brooks, and B. S. Middleditch, Anal. Chem, 44, 30-37, 1972. J-P. Thenot and E. C. Homing, Anal. Letters, 5, 21-33, 1972. EM. Chambaz, G. Defaye, and C. Madani, Anal. Chem. 45, 1090-1098, 1973. C. J. W. Brooks, W. J. Cole, T. D. V. Lawrie, J. MacLachlan, J. H. Borthwick, and G. M. Barrett, J. Steroid Biochem, 19, 189-201, 1983. R. W. Kelly, Steroids, I3, 507, 1969. 27 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 28 M. Ishibashi, M Itoh, K Yamashita, H Miyazaki, and H Nakata, Chem. Pharm. Bull., 34, 3298-3305, 1986. C. F. Poole, W. F. Sye, S. Singhawangcha, and E. D. Morgan, Lipids, 15, 734- 744, 1980. S. J. Gaskell, C. J. Porter, and B. W. Green, Biomed. Mass Spectrom, 12, 139- 141, 1985. V. J. Gould, A. O. Turkes, and S. J. Gaskell, J. Steroid Biochem, 24, 563-567, 1986. ’ D. F. Hunt, G.C. Stafford, Jr., F. W. Crow, and J. W. Russell, Anal. Chem, 48, 2098-2105, 1976. H. Budzikiewicz, Angew. Chem. Int. Ed. Engl., 20, 624-637, 1981. C. E. Klots, "Energy Deposition Processes" in Fundamental Processes in Radiation Chemistry, P. Ausloos, ed., John Wiley and Sons, New York, p. 40, 1968. R. C. Dougherty and C. R. Weisenberger, J. Am. Chem. Soc., 90, 6570, 1968. D. F. Hunt and F. W. Crow, Anal. Chem, 50, 1781-1784, 1978. I. A. Blair, S. E. Barrow, K. A. Waddell, and P. J. Lewis, Prostaglandins, 23, 579-589, 1983. K. F. Faull and 0. Beck, "Quantification of Neurotransmitters and Related Compounds by Electron Capture Chemical Ionization GC-MS" in Mass Spectrometry in Biomedical Research, 8. J. Gaskell, ed., John Wiley and Sons Ltd., Chichester, pp. 403-421, 1986. W. A. Garland and M. P. Barbalas, "Determination of Drugs in Biological Matrices Using Gas Chromatography-Electron Capture Negative Chemical Ionization Mass Spectrometry" in Mass Spectrometry in Biomedical Research, S. J. Gaskell, ed., John Wiley and Sons Ltd., Chichester, pp. 379-401, 1986. D. Stock] and H. Budzikiewicz, Org. Mass Spectrom, 17, 470-474, 1982. T. G. Heath, D. A. Gage, J. Zeevart, and J. T. Watson, Proceedings of the 37th Annual Conference on Mass Spectrometry and Allied Topics, Miami Beach, FL, May 1989. D. B. Kassel, K. A. Kayganich, J. T. Watson, and J. Allison, Anal. Chem, 60, 911-917, 1988. L. J. Sears, J. A. Campbell, E. P. Grimsrud, Biomed. Environ. Mass Spectrom, 14, 401-415, 1987. V. Burger and P. J. Pretorius, J. Chromatogr. Sci, 25, 118-121, 1987. E. A. Stemmler and R. A. Hites, Biomed. Environ. Mass Spectrom, 15, 659-667, 1988. 37. 38. 39. 40. A 41. 42. 43. 45. 46. 47. 48. 49. 50. 51. 52. 2 9 E. A. Stemmler, R. A. Hites, B. Arbogast, W. L. Budde, M. L. Deinzer, R. C. Dougherty, J. W. Eichelberger, R. L. Foltz, C. Grimm, E. P. Grimsrud, C. Sakashita, and L. J. Sears, Anal. Chem, 60, 781-787, 1988. M. Dawson, C. M. McGee, P. M. Brooks, J. H. Vine, and T. R. Watson, Biomed. Environ. Mass Spectrom, 17, 205-211, 1988. R. c. Dougherty, Anal. Chem, 53, 625A-636A, 1981. B. J. Miwa and w. A. Garland, Anal. Chem, 53, 793-797, 1981. CF. Poole and A. Zlatkis, Anal. Chem, 52, 1002A-1016A, 1980. J. Goto, K Watanabe, H. Miura, T. Nambara, and T. Iida, J. Chromatrogr, 388, 379-387, 1987. J. E. Lovelock, P. G. Simmonds, and W. J. A. Vandenheuvel, Nature, 197, 249- 251, 1963. R. A. Davis, K. T. Koshy, M. L. Ogilvie, G. L. Stahl, and W. A. Callahan, J. Assoc. 01?. Anal. Chem, 56, 103, 1973. E. D. Morgan and C. F. Poole, J. Chromatogr, 116, 333, 1976. F. H. de Jong and H. J. van der Molen, J. Endocr, 53, 461-474, 1972. G. R. Her and J. T. Watson, Biomed. Environ. Mass Spectrom, 13, 57-63, 1986. G. R. Her and J. T. Watson, Anal. Biochem, 151, 292-298, 1985. R. C. Dougherty, J.’ Dalton, F. J. Biros, Org. Mass Spectrom, 6, 1171, 1972. F. H. Field, Proceedings of the 28th Annual Conference on Mass Spectrometry and Allied Topics, New York, NY, pp. 2-13, May 1980. L. G. Christopherou, Chem. Rev, 76,409, 1976. J. G. Dillard, Chem. Rev, 73, 589-643, 1973. CHAPTER 2. DEVELOPMENT OF OPTIMAL CONDITIONS FOR ISOLATION AND CHEMICAL OXIDATION OF DEXAMETHAS ONE 1. Introduction As described in the previous chapter, dexamethasone is a synthetic, corticosteroid, anti-inflammatory drug that, upon oxidation, yields a highly electrophilic 11,17-keto analog. Besides its use as an anti-inflammatory drug, dexamethasone has been found to aid in the diagnosis of Cushing’s Syndrome, a hypercortisolemic condition. Upon administration of dexamethasone, plasma cortisol levels of healthy people are suppressed for a 24-hour period. In patients having Cushing’s Syndrome the administration of dexamethasone has no effect upon the plasma cortisol levels. The results of this test, coined the dexamethasone suppression test, has served as a basis for diagnosing Cushing’s Syndrome for many years. Recently (within the last decade) the dexamethasone suppression test (DST) has been proposed as a tool for physicians and psychiatrists in the diagnosis of endogenously depressed psychiatric patients. The DST has been shown to aid in differentiation between melancholia (endogenous depression) and depression caused by neurotic, reactive, or characterological factors (1). Because the treatments for these types of depression differ, it is important to be able to obtain a proper diagnosis. The DST typically consists of administration of 1 mg of dexamethasone with subsequent monitoring of plasma cortisol levels at 8, 16, and 24 hours following administration. In non-endogenously depressed subjects, cortisol levels are suppressed for a 24-hour period, whereas in endogenously depressed patients, the plasma cortisol levels ’escape’ to pre- drug levels within three to four hours. Some recent studies have indicated that the bioavailability of dexamethasone may be a critical factor for determining the results of the DST (2). Therefore, it appears necessary to measure dexamethasone plasma levels 30 3 1 along with cortisol levels in order to appropriately evaluate DST results. Typical dexamethasone concentrations in plasma of patients undergoing the DST diminish from 4 ng/ml to 0.1 ng/ml during the 24-hour period of the test. Due to the low concentrations of dexamethasone and the complexity of the plasma sample matrix, sensitive and specific methodology is needed for accurate determination of the concentration of dexamethasone in plasma. There are adequately sensitive radioirnmunoassays available to measure these levels of dexamethasone in plasma (3,4); however, the possibility of cross-reactivity of the antibody with endogenous compounds as well as dexamethasone metabolites may make RIA results dubious. HPLC methods employing UV detection are also available for the determination of dexamethasone in plasma. The sensitivity of these methods are limited, however, due to interferences from co-extracted endogenous material. HPLC methods are suitable for measuring plasma dexamethasone concentrations no lower than 3 to 5 ng/ml (5). Mass spectrometric methods have also been reported. Some of these reported MS methods are based on the formation of the tetra-trimethylsilyl ether (TMS) derivative of dexamethasone with detection by gas chromatography-mass spectrometry with either chemical ionization or electron impact ionization (6,7). Using selected ion monitoring GC-EI/MS Kasuya et a1 (6) were able to detect an injection of 100 pg of dexamethasone (S/N = 2.5) from 1 ml of plasma spiked to a concentration of 300 pg/ml. Extensive isolation procedures, consisting of extraction by C18-bonded phase reverse-phase cartridge followed by normal phase HPLC purification, were necessary to eliminate the plasma matrix interferences. A method based on direct liquid introduction microbore HPLC-MS has been reported for the detection of betamethasone in equine urine (9). This drug is identical to dexamethasone except in the sterochemistry of the 16-methy1 group. The reported sensitivity of 5 ng/ml would not be adequate for the determination of dexamethasone in the plasma of patients undergoing the DST. In addition to the measurement of plasma dexamethasone concentrations, the 32 measurement of 6B-hydroxydexamethasone, the major metabolite of dexamethasone (8), may also be desirable since there is some indication that 6B-hydroxydexamethasone is bioactive (10). A goal of this research is to apply oxidation methodology for the detection of dexamethasone and 6B-hydroxydexamethasone in plasma at levels expected to be present in patients undergoing the DST. This chapter describes the optimization of plasma extraction, extract oxidation, removal of excess oxidation reagent for the detection of dexamethasone in plasma. Some preliminary work for the application of this methodology to the detection of 6B-hydroxydexamethasone is also described. II. Method Development The oxidation method consists of four parts; plasma extraction, extract oxidation, removal of excess oxidation reagent, and finally ECNI-MS analysis. The oxidation of dexamethasone to its 11,17-keto analog is shown in Figure 2.1. This is the reaction that serves as the basis for the ECNI-MS assay of dexamethasone. The following sections describe the development and optimization of experimental conditions and briefly discuss the results for isolation procedures and chemical oxidation methodology for dexamethasone and 6B-hydroxydexamethasone. A. Extraction of Dexamethasone and 6B-Hydroxydexamethasone from Plasma 1. Solvent Extraction vs. Solid Phase Extraction Solvent extraction was once the most extensively used means by which to isolate steroids from physiological fluids such as plasma and urine. The development of 1 dexamethasone Figure 2.1. Oxidative conversion dexamethasone. 33 Oxidation of dexamethasone [0 11,17-keto 34 reverse-phase, solid-phase extraction technology has provided a more rapid means for the isolation of steroids from these fluids. Near quantitative recoveries of steroids from urine through the use of C18 Sep-Pak cartridges have been reported (13). In addition to high recovery, solid-phase extraction technology allows the possibility of isolating the analyte from many interfering substances within the sample matrix. Aliquots of "spiked" plasma were extracted by solvent extraction and C18 solid- phase extraction to compare the efficiency of each method for the recovery of dexamethasone, 13C6-2H3-dexamethasone (an isotopically labelled analog of dexamethasone used as an internal standard), and 6B-hydroxydexamethasone from human plasma. The plasma samples were prepared by adding dexamethasone (16.7 ng), 6B-hydroxydexamethasone (19.4 ng), and 13C6-2H3-dexamethasone (35 ng) to ten ml of blood bank plasma. Aliquots of 2 ml were extracted by each method. Figure 2.2 illustrates the impact of each of the extraction procedures on the GC-ECNI/MS response obtained by analysis of these test samples. From Figure 2.2 it is apparent that the recovery of the dexamethasone is much more efficient by solid-phase extraction than by solvent extraction with ethyl acetate. However, the intensity of the ion current from co—extracted matrix components is about the same by solvent extraction and solid-phase extraction as can be seen by comparing the ion current intensity numbers on the right border of Figure 2.2a and Figure 2.2b. By varying the solvent strength of the column washes and the eluate, it would probably be possible to eliminate some of these co-extracted compounds. This option would not be possible in the case of solvent extraction. The solvent extraction method consisted of extraction of the plasma three times with equal volumes of ethyl acetate. The combined ethyl acetate extracts were then dried with anhydrous sodium sulfate. The extract was evaporated under N2, reconstituted in lml of CH2C12, and oxidized with pyridinium chlorochromate according to procedures %100- 80' 40‘ 20' 35 -*E+04 4.706 l A 1'1‘11 10: 00 uI-Ivlu'ulvlv'w 7:00 8:00 1' '1-I . . . . 9:00 11:00 12:00 .*E+04 4.072 0 T Figure 2.2. ujulviv 'vlefil 7: 00 8: ul‘lv1v 1., . . 00 9:00 10:00 Y'iIvlvl ' 1 11:00 12:00 time (min) Reconstructed selected ion current profiles from GC-ECNI/MS analysis of oxidized dexamethasone, 13C6-2H3-dexamethasone, and 6B-hydroxydexamethasone isolated from plasma by a) solvent extraction and b) C18 solid-phase extraction. Arrow indicates retention time of oxidized dexamethasone. 3 6 described in IIB. of this chapter. Conditions for analysis of the sample by GC-ECNI/MS with SIM are described in Chapter 3. For the solid phase extraction, two milliliters of the spiked plasma was passed through a 200 mg C18 solid phase extraction column that had been previously washed with 5 ml of 8M urea, 5 ml of methanol, and 10 ml of distilled water. The purpose of the urea wash is to derivatize any free silanol groups on the column packing. The methanol wash elutes any contaminants sticking to the C18 packing. The water wash removes the methanol and prepares the column to receive the sample. Following the application of the plasma sample, the column was washed with 4 ml of distilled water to remove polar components of the plasma sample. To elute the dexamethasone and 6B- hydroxydexamethasone, four milliliters of 70% methanol-30% hexane was applied to the column and the eluate collected. The eluate was evaporated under N2 and reconstituted in 1 ml CH2C12 for oxidation followed by GC-ECNI/MS analysis. 2. Solid—Phase, Reverse-Phase (C18) Extraction Two different solid-phase extraction methods were investigated for the isolation of dexamethasone and 6B-hydroxydexamethasone in human plasma. The method of Plezia and Berens (14) involved conditioning a C18 SPE column in tandem with a silica SPE column with 2 ml methanol and then 2 ml water. After passing the sample through the extraction system, the silica phase was removed and the C18 column was washed with 3 ml 95% aqueous methanol and 4 ml of CHC13. The dexamethasone was eluted with 2 ml of methanol-CHC13 (3:1). The authors reported 72% extraction recovery. Also, many of the matrix components that would interfere with their analysis of dexamethasone by HPLC were removed by the washes. A preliminary experiment using this method indicated that both dexamethasone and 6B-hydroxydexamethasone were recovered from a real plasma sample taken from a 37 DST patient. Analysis of the oxidized plasma extract by GC-ECNI-MS with SIM revealed peaks corresponding to the masses and retention times of oxidized dexamethasone and oxidized 6B-hydroxydexamethasone. Recovery of dexamethasone by this solid phase extraction method was not determined rigorously. The estimated recovery of dexamethasone was 65%. This estimate was based on comparison of the GC-FID peak area of the 17-keto decomposition peak, resulting from the injection of a known amount of dexamethasone, with the peak area of the 17-keto decomposition peak resulting from a known amount of dexamethasone that had been extracted according to the Plezia and Berens method. This recovery is reasonably comparable to that found by the authors. Because the Plezia and Berens solid phase extraction method was deve10ped as purification step to be used prior to HPLC analysis, the elimination of interfering substances was of utmost importance to these authors, as the selectivity of the UV detector of the HPLC is limited in the analysis of complex samples. In fact, the purification of the dexamethasone from interfering matrix components was optimized at the expense of a quantitative recovery of dexamethasone. Because ECNI mass spectrometry with SIM is a highly selective means of detection, and because oxidation selectively enhances the response of dexamethasone within the extracted plasma matrix, removal of matrix components is not as necessary for ECNI-MS analysis. Instead, a quantitative recovery of dexamethasone is more desirable to allow the lowest possible detection limit. Therefore, less selective solid phase extraction methodology that would maximize the recovery of dexamethasone and 6B-hydroxydexamethasone was investigated. A solid phase extraction method for dexamethasone was developed by Clinton Kilts and Jim Ritchie at Duke University (10). As part of our collaboraton with the Duke laboratory we received the protocol for their solid phase extraction method. The method consisted of the following steps: 3 8 1) precondition a 200mg C18 SPE column with a) 5 ml 8M urea b) 5 ml methanol c) 10 ml distilled water 2) pass sample through column, 1-2 drops/sec 3) wash column with a) 4 ml distilled water b) 4 ml 38% methanol-water 4) elute dexamethasone with 4 ml 70% methanol-30% hexane Through the use of a radiolabelled analog of dexamethasone, Kilts and Ritchie determined the recovery of dexamethasone by this method to be 85%. Preliminary experiments with this extraction method were satisfactory for the recovery of dexamethasone. However, the recovery of 6B-hydroxydexamethasone was very poor. Since 6B-hydroxydexamethasone is more polar than dexamethasone, this compound was probably eluted by the the 38% methanol-water wash, whereas the more non-polar dexamethasone remained on the non-polar reverse-phase column. 3. Influence of Composition of Column Wash (CH3OHszO) on Recovery of 6B- Hydroxydexamethasone In order to find conditions that would improve the recovery of 6B- hydroxydexamethasone, an experiment was done in which the percentage of methanol in the methanol-water wash was decreased. Four 0.5 ml aliquots of plasma, each spiked with 230 ng of 6B-hydroxydexamethasone and 420 ng of dexamethasone, were loaded onto four, 200 mg, C18 SPE columns. The extraction procedure outlined above was followed except that the percentage of methanol in the methaonol-water wash was different for each of the four extractions. In order to estimate the percent recoveries of 3 9 dexamethasone and 6B-hydroxydexamethasone by each of the extractions, the eluates were reconstituted in 200 111 of CH2C12 and analyzed by normal phase HPLC. The peak areas obtained for dexamethasone and 6B-hydroxydexamethasone were compared to the peak areas obtained from a standard which consisted of 230 11g 613- hydroxydexamethasone and 420 ng dexamethasone in 200 111 of CH2C12. Because the eluates could be analyzed directly, HPLC provided a faster and simpler means to determine the percent recoveries than would oxidation followed by GC-ECNI/MS analysis. The results of this experiment are shown in Table 2.1. Conditions for the HPLC separation were as follows: partisil silica gel column (25cm x 4.6 mm, 10 11 particle size), 7% ethanol-33% CH2C12-60% hexane mobile phase, 1.8 ml/min = flow rate, injection volume = 20 111, wavelength = 240 nm, 0.1 absorbance units full scale. Table 2.1 Extraction recoveries ' of dexamethasone (dex) and 6B- hydroxydexamethasone (6B-OH dex) vs. percentage of methanol in solid phase extraction wash. % methanol % recovered 6 -OH dex % recovered dex 38 40 0.0 25 1 13 25 15 104 98 5 98 84 Most of the dexamethasone and 6B-hydroxydexamethasone was recovered if the methanol-water wash was 15% or less in methanol. The quantities to be recovered from real plasma samples will be on the order of 100 times less. Also, when extracting real plasma samples, the sample volume will be two to four times larger than those used in this experiment. Therefore, it is possible that a 15% methanol-water wash may not represent the optimum for analysis of realistic samples. However, it is likely to be close to the optimum condition. Therefore, the 15% methanol-water wash will be used for all subsequent extractions of dexamethasone and 6B-hydroxydexamethasone from plasma. 40 B. Oxidation Previous work in this laboratory indicated a yield of about 30% 11,17-keto dexamethasone upon 3— to 5-hour oxidation of dexamethasone with chromium trioxide- pyridine complex (pyr2CrO3) in pyridine (11). After evaporation of the pyridine, the product(s) was isolated from the reaction mixture by filtration through a small sephadex LH-20 column (3cm x 0.5 cm) with ethyl acetate. The yield could be increased slightly, to about 45% through the use of NaBiO3, a reagent that is specific for oxidation of 1,2- diols or 1,2—hydroxy ketones, to cleave off the C-17 side-chain followed by pyr2CrO3 oxidation to oxidize the ll-hydroxyl. However, this involves a two-step process which requires more labor and could lead to losses of analyte during transfer. A goal of this research is to improve the sensitivity of the assay for dexamethasone in order to be able to detect 100 pg/ml concentrations of dexamethasone in 1 to 2 ml plasma samples. Improving the oxidation yield of the 11,17-keto dexamethasone ought to improve the sensitivity of the assay. However, another goal is to avoid complicating the procedure so as to provide a simple approach for the analysis of plasma for dexamethasone. Therefore, rather than use a two-step oxidation procedure, several parameters including choice of reagent, reaction time, reaction solvent, heat, and the presence of a base were investigated as means for increasing the oxidative yield of 11,17-keto dexamethasone. 1. Choice of Oxidation Reagent The requirements for a reagent to be used for the oxidation of dexamethasone are: (1), the reagent is capable of producing a high yield of the 11,17-keto analog of dexamethasone in a reasonable amount of time; (2), losses due to over-oxidation (beyond the ketone) are minimal; (3), separation of the oxidation reagent from the dexamethasone 4 l product(s) is easily achieved; and (4), the reaction involves only one step. There are a wide variety of chemical oxidation reagents available for the oxidation of organic compounds. Chromium trioxide is a very strong oxidizing agent that has been used for oxidation of dexamethasone to its 11,17-keto analog (12). However, this reagent is capable of oxidizing a hydroxy group all the way to a carboxyl group and may also react with double bonds and other functionalitie‘s. This would make it difficult to convert dexamethasone to only one product, namely 11,17-keto dexamethasone. Also, the stronger the oxidation reagent, the more oxidized the sample matrix will become. For instance, an alkene can be converted to an a,B-unsaturated ketone through the oxidation of the allylic C-H bond by any of several chromium (VI) oxidants (15). Increased oxidation of the matrix components could increase the electron capture response of the matrix, thereby increasing the probability of interference from the sample matrix. For these reasons, milder oxidation reagents were investigated. The derivatives of chromium oxides that were investigated included pyridinium dichromate (PDC), pyridinium chlorochromate (PCC), poly(4-viny1pyridinium dichromate) (PVPDC) which is an immobilized reagent, and pyrzCrO3 complex. The structures of these reagents are shown in Figure 2.3. An initial experiment was done to compare the yields of 11,17-keto dexamethasone that could be obtained with PCC and PVPDC with those obtained using pyr2CrO3 under the conditions used previously in this laboratory. These results are shown in Table 2.2. Dexamethasone (260 11g) was placed into three silanized vials. (All glassware used in these procedures was silanized to prevent adsorptive losses of dexamethasone or its oxidized products.) For the PCC oxidation, the dexamethasone was dissolved in 1 ml CH2C12; excess PCC, that had been finely ground with celite, was then added to the vial. The celite provides a surface for the reaction to occur and also aids the ~ AWN 42 O Oicrfo CN+—Cr—O‘ [CN] Cro, X X — (ll) ‘ 2 X=OH 5 6 X=OAc X=Cl X=O—t—Bu :1 I crcro; a 0,0; M C100,- \N ‘N FIN- H 2 8 4:112 9 Figure 2.3. Structures of some chromium (VI) oxidants; chromic acid (1), chromyl acetate (2), chromyl chloride (3), tert—butyl chromate (4), chromium trioxide-pyridine complex (5), dipyridine chromium (VI) oxide (6), pyridinium chlorochromate (7), pyridinium dichromate (8), and 2,2’-bipyridinium chlorochromate (9). Adapted from reference (15). 4 3 removal of the reagent by adsorbing the reduced chromium salts. Used without celite, PCC is reported to deposit a black insoluble tar as the reagent becomes reduced (16). For the pyr2CrO3 oxidation, the reagent was added as a solution in pyridine (0.5 ml of 7.5 mg/ml). To remove the excess reagent and isolate the products, the solvents were evaporated and the residues reconstituted in 0.5 ml ethyl acetate. These were then transferred to short Sephadex LH—20 columns. About 3 ml of ethyl acetate eluate was collected. For the PVPDC oxidation the reagent was soaked in distilled water overnight and filtered in the morning to remove the water and the non-immobilized chromium salts. The dexamethasone was dissolved in 1 ml of cyclohexane followed by addition of the wet PVPDC. To remove the PVPDC, the reaction mixture was filtered through a glass wool plug in a pasteur pipette. Other details about the oxidations are included in Table 2.2. Table 2.2 Relative yields of 11,17-keto dexamethasone obtained with three different oxidation reagents. reagent rel. %yld reaction conditions pyrzCrO3 5.1 stirred, room temp., 1 2/3 hr., pyridine PCC 66 stirred, room temp, 1 1/2 hr., CH2C12 PVPDC 100 70C, 21 hrs., cyclohexane Oxidation of alcohols with Cr(VI) oxidants usually proceeds with the formation of a chromate ester followed by abstraction of a hydrogen from the reaction substrate and the elimination of the reduced chromate. This mechansim is shown in Figure 2.4. The rate limiting step is apparently the abstaction of the hydrogen from the substrate. Tertiary alcohols have no hydrogen atom on the same carbon atom to which the hydroxy group is attached; therefore, these compounds are relatively inert to oxidation with Cr(VI) oxidants. The cleavage of the C-17 sidechain of dexamethasone to form 44 (It) . \ /o’—“cu:r—orr \ (I? / \P{’ HO/ \OH Figure 2.4. Mechanism for the oxidation of alcohols with chromium (V I) oxidants. 4 5 the 17-ketone involves the oxidation of a tertiary hydroxyl group. Therefore, a more complicated oxidation mechanism may be involved in this reaction. A mechanism suggested for the oxidative cleavage of 1,2-diols is illustrated in Figure 2.5a. Chromium (VI) oxidants are capable of cleaving the C-C bond of a diol which results in two ketone products. The sidechain of dexamethasone does not contain a 1,2-diol moiety. However, the ketol present on the sidechain may be capable of the tautomerization shown in Figure 2.5b that would provide a 1,2-diol for oxidative cleavage by Cr(VI) which would result in the production of 17-keto dexamethasone. This mechanism may not. be operative in the production of 17-keto dexamethasone, however, no definite mechanism for the oxidation of these types of corticosteroid sidechains by chromium reagents has been found in literature as this type of destructive reaction is not normally considered to be synthetically useful. 45 the 17-ketone involves the oxidation of a tertiary hydroxyl group. Therefore, a more complicated oxidation mechanism may be involved in this reaction. A mechanism suggested for the oxidative cleavage of 1,2-diols is illustrated in Figure 2.5a. Chromium (VI) oxidants are capable of cleaving the C—C bond of a diol which results in two ketone products. The sidechain of dexamethasone does not contain a 1,2-diol moiety. However, the ketol present on the sidechain may be capable of the tautomerization shown in Figure 2.5b that would provide a 1,2-diol for oxidative cleavage by Cr(VI) which would result in the production of 17-keto dexamethasone. This mechanism may not. be operative in the production of 17-keto dexamethasone, however, no definite mechanism for the oxidation of these types of corticosteroid sidechains by chromium reagents has been found in literature as this type of destructive reaction is not normally considered to be synthetically useful. 46 1'1 R I R—C—OH R—C—O ,o R 1 +H,cro,——- I \Cr’ -—»2 \C==O+Cr(lV) R—C—OH R—C—O/ ‘0 / 1 1 R R R OH 0 OH O --OH --OH Cr(VI) - ~CH3 3:: --CH3 —» --CH3 Figure 2.5. a) Mechanism for the oxidative cleavage of the C-C bond of a 1,2- diol and b) proposed mechanism for the oxidative cleavage of the dexamethaspne C-17 sidechain. 47 Each reaction produced a single product as indicated by a single peak by GC with flame ionization detection (FID). The peak from each reaction eluted from the GC column at the time predicted by the retention index of 11,17-keto dexamethasone. No other peaks were observed in the chromatograms. The results in Table 2.2 indicate that the yields achievable with PCC and PVPDC are likely to be much higher than those with pyr2CrO3 under these oxidation conditions. One possible reason for the relatively poor yield achieved with pyr2CrO3 is that the reaction was conducted in pyridine. Reaction rates with pyr2CrO3 are reported to be higher in CH2C12 than in pyridine (17). This is probably because the reagent is twice as soluble in CH2C12 as it is in pyridine. Better yields might have been achieved with pyr2CrO3 if CH2C12 had been used. However, due to the relatively high yields achieved with the other reagents, further study of py12CrO3 was abandoned and studies of PCC and PVPDC were pursued. Since 2/3 of the yield achieved with PVPDC was achieved in 1/10 of the time with PCC, PCC seemed the better candidate for the most efficient oxidizing agent. PVPDC, however, has an advantage of very simple clean-up because the reagent is immobilized on polyvinyl beads. To remove the reagent, the reaction mixture is simply filtered through a glass wool plug, and the polyvinyl beads are collected. Because of this advantage an experiment was done to determine if high yields could be achieved with PVPDC at shorter reaction times. These results are shown in Table 2.3. Table 2.3 Reaction time vs. yield of 11,17-keto dexamethasone by PVPDC oxidation of dexamethasone at 70°C, in cyclohexane. reaction time rel. yield ca. %yield 1 hr 0.02 1.7 3 hr 0.23 19 4.5 hr 0.36 30 24 hr 1.00 83 4.5 , CH2C12-acetone 0.18 15 4 8 Percent yields of 11,17 -keto dexamethasone were approximated by comparison of TIC peak areas to a known amount of 1,4-androstadiene-3,11,17-one obtained by GC- EI/MS. Unfortunately, the results in Table 2.3 cannot be directly compared to those in Table 2.2. However, if one assumes that the yields of the 21-hour PVPDC oxidation in Table 2.2 and the 24-hour PVPDC oxidation in Table 2.3 are equal, then as an approximation it appears that the yield of 11,17-keto dexamethasone by PCC oxidation for 1.5 hours is still about two times higher than that achieved with PVPDC at a reaction time of 4.5 hours. The yields at times greater than 4.5 hours were not investigated since reaction times much greater than 5 hours would not allow the whole sample preparation (extraction, oxidation, and removal of excess reagent) to be completed in a regular working day. Instead, other factors were investigated which might increase the yields available with PVPDC. Because dexamethasone is only sparingly soluble in cyclohexane, CH2C12- acetone was added to the reaction mixture to enhance the solubility of dexamethasone. Rather than increase the yield, this decreased the yield of the reaction as can be seen in the last entry of Table 2.3. Perhaps the more polar solvent solvated the dexamethasone more strongly, leading to a slower rate of oxidation. Or, the conformation of the polyvinyl polymer beads was changed in a way that slowed the reaction. When a reaction with wet PVPDC is carried out in cylcohexane, the PVPDC sits on the bottom of the vial inside the bead of water. When the reaction was carried out with CH2C12- acetone, the reaction mixture was a cloudy emulsion. It is recommended in the original journal article (18) that the PVPDC be washed with water to remove loose chromium salts and then soaked in water and filtered, but not dried, prior to use. The dry reagent was found to be inactive. Perhaps the use of polar solvents displaces from the PVPDC some of the water necessary for oxidation. Because some water is necessary for PVPDC activity, another experiment was designed to investigate whether the amount of water present during the reaction would 4 9 affect the yield of 11,17-keto dexamethasone achieved with PVPDC. The effects of heating and stirring upon the yield also were compared in this experiment. These results are shown in Table 2.4. Table 2.4 Relative yield of 11,17-keto dexamethasone by PVPDC oxidation vs. presence of excess H20 and heating and stirring. 26 ttg of dexamethasone was used for each oxidation. time reagent conditions relative field 3 hr PCC stirred 100 3 hr PVPDC heated 40 3 hr PVPDC heated, + H20 46 4.5 PVPDC stirred 33 4.5 PVPDC heated 84 4.5 PVPDC heated and stirred 74 From the results in Table 2.4, a number of observations may be made. First, stirring the PVPDC reaction does not increase the oxidation yield. Second, heating is neccessary to promote a relatively high yield with PVPDC. Third, the presence of extra water in the PVPDC reaction may slightly increase the yield. However, the 6% difference between the yields achieved by PVPDC with and without excess water may not be significant because of experimental error; each oxidation was done only once. Finally, the yield achieved with the 3 hour PCC oxidation is still higher than any of the yields achieved using PVPDC under all the different conditions. Therefore, further oxidations with PVPDC were abandoned and effort was concentrated on maximizing the yield by PCC oxidation. A related reagent, PDC, also was investigated. PDC is a neutral reagent, unlike PCC which is slightly acidic. Therefore, PDC is used especially for oxidation of compounds contained acid sensitive groups. Since dexamethasone contains an 01,13- unstaurated ketone, it is possible that it would be acid sensitive. Also, because PDC is more polar than PCC, it cannot be eluted from a silica column by ethyl acetate during 4 9 affect the yield of 11,17-keto dexamethasone achieved with PVPDC. The effects of heating and stirring upon the yield also were compared in this experiment. These results are shown in Table 2.4. Table 2.4 Relative yield of 11,17-keto dexamethasone by PVPDC oxidation vs. presence of excess H20 and heating and stirring. 26 pg of dexamethasone was used for each oxidation. — time reagent conditions relative yield 3 hr PCC stirred 100 3 hr PVPDC heated 40 3 hr PVPDC heated, + H20 46 4.5 PVPDC stirred 33 4.5 PVPDC heated 84 4.5 PVPDC heated and stirred 74 From the results in Table 2.4, a number of observations may be made. First, stirring the PVPDC reaction does not increase the oxidation yield. Second, heating is neccessary to promote a relatively high yield with PVPDC. Third, the presence of extra water in the PVPDC reaction may slightly increase the yield. However, the 6% difference between the yields achieved by PVPDC with and without excess water may not be significant because of experimental error; each oxidation was done only once. Finally, the yield achieved with the 3 hour PCC oxidation is still higher than any of the yields achieved using PVPDC under all the different conditions. Therefore, further oxidations with PVPDC were abandoned and effort was concentrated on maximizing the yield by PCC oxidation. A related reagent, PDC, also was investigated. PDC is a neutral reagent, unlike PCC which is slightly acidic. Therefore, PDC is used especially for oxidation of compounds contained acid sensitive groups. Since dexamethasone contains an 01,13- unstaurated ketone, it is possible that it would be acid sensitive. Also, because PDC is more polar than PCC, it cannot be eluted from a silica column by ethyl acetate during 5 O removal of the reagent following the oxidation. PCC begins to elute with ethyl acetate and sometimes may not be completely separated from the oxidized dexamethasone. If, after optimization of the silica clean-up step, it is found that the PCC eluted with the oxidized dexamethasone, then PDC could possibly be used as an alternative to PCC. The results from a preliminary study using PDC are shown in Table 2.5. For these reactions excess PDC finely ground up with celite was added to 50 ug of dexamethasone that had been dissolved in 1 ml CH2C12. The reactions were stirred at room temperature. Excess PDC was removed by passing reaction mixture through a 500 mg silica column. Ethyl acetate was added to elute the oxidized dexamethasone. Relative yields are reported plus or minus (i) one standard deviation calculated from the GC-FID peak areas of at least three separate injections. Table 2.5 Relative yield of 11,17-keto dexamethasone by PDC oxidation vs. time. reaction time rel. yield 2 hrs. 77 i 8 3 hrs. 78 5 hrs. 100 :1: 1 (ca. yld. 50%) 7 hrs. 96 i 3 9 hrs. 78 i- 10 The yield achieved with the PDC oxidation is maximized between 5 and 7 hours. The yield appears to decrease if the reaction is allowed to continue for 9 hours. Over- oxidation could be starting to occur at the longer reaction times. The yield of the 5-hour reaction was estimated to be 50% by comparing the peak area of the 11,17-keto dexamethasone to the peak area of a known quantity of 170t- methyl testosterone. Since Nor-methyl testosterone contains the same number of carbon atoms as 11,17-keto dexamethasone it was initially assumed that the FID responses per nanogram of the two steroids would be essentially equal. However, the FID response of a compound is not simply proportional to the number of carbon atoms within a molecule, 51 but to the number of oxidizable carbon atoms within the molecule (19). Because 11,17- keto dexamethasone is more highly oxidized than Nor-methyl testosterone, the response per nanogram of 11,17-keto dexamethasone is less than the response per nanogram of Nor-methyl testosterone. Therefore, yields of 11,17-keto dexamethasone calculated by comparison to the response of a known amount of 1701-methyl testosterone will be atificially low because a greater amount of 11,17-keto dexamethasone must be injected to achieve the same peak area as produced by the given amount of Nor-methyl testosterone. In order to determine the yield of 11,17-keto dexamethasone accurately from any oxidation, it was necessary to Obtain a primary standard of 11,17-keto dexamethasone which could be weighed on an analytical balance. Therefore, a large quantity of dexamethasone was oxidized and recrystallized. The reaction was scaled up in order to oxidize 18 mg of dexamethasone. The reaction was followed by thin layer chromatography (TLC). After 1.5 hours only one spot was apparent. This spot had the same Rf value as an authentic standard of 11,17-keto dexamethasone. After removal of the excess oxidation reagent on a silica column and recrystallization of the product from aqueous acetone, a quantity of the product was weighed. GC-FID analysis of a solution of the product resulted in one peak having the retention time of authentic 11,17-keto dexamethasone. Later, GC-EI/MS confirmed the identity of the product as 11,17-keto dexamethasone. Serial dilutions of this solution were made in order to construct standard curves for the determination of unknown concentrations of 11,17-keto dexamethasone by GC with FID and ECD. 2. Optimization of PCC Oxidation Conditions In order to improve the yield of 11,17-keto dexamethasone achieved by PCC oxidation, the effects of heating and the presence of a base (either sodium acetate or pyridine) were studied. Sodium acetate has been used to buffer the acid liberated during 52 the PCC oxidation. A small percentage of pyridine in CH2C12 could also be used as a buffer. However, the selectivity of the reaction has been observed to be altered with 2% pyridine. Sodium acetate has no effect on the selectivity of PCC oxidation. Tables 2.6 and 2.7 summarize the results of these studies of the effect of heat, sodium acetate, and pyridine upon the yield of 11,17-keto dexamethasone by PCC oxidation. All of the oxidations were carried out for 2 1/4 hours in 1.0 ml of CH2C12, with an excess of finely ground pyridinium chlorochromate (PCC)/Celite mixture. Fifty micrograms of dexamethasone were used in each oxidation. Removal of the excess oxidation reagent and recovery of the oxidation products were done by passing the reaction mixture through a 500 mg silica solid phase extraction (SPE) column. The products were eluted from the column with 2.8 ml of ethyl acetate. Table 2.6 The effect of a base on the yield of 11,17-keto dexamethasone. vial no. treatment %yield 1a N aOAc 54 lb NaOAc 55 2 1% pyridine 59 3 none 27 The percent yields were calculated by comparison of the GC-FID peak areas to a standard curve constructed with the serial dilutions of the known concentration of the recrystallized 11,17-keto dexamethasone. From comparison of the percent yields for each reaction in Table 2.6 it is apparent that the presence of a base in the oxidation enhances the rate at which the 11,17-keto dexamethasone is produced. Also to be noted from Table 2.6 is the closeness of the results for the reactions which contained a base (vials 1a and 1b, which were duplicates, and vial 2). Since the relative standard deviation (RSD) for GC analysis is 2.4%, it is difficult to determine whether there is any difference between the yields when pyridine is used as the base or when NaOAc is used. 53 The closeness of the results for the duplicate reactions, 1a and 1b, is encouraging with regard to the reproducibility of the oxidation reaction. The difference between the percent yields for these reactions is within the variation to be expected due to GC injection and integration errors. , It should also be mentioned that the percent yield calculated for the reaction in which no base was used may be artificially low due to losses of sample during transfer steps. To prevent this kind of ambiguity an internal standard, 4-androstene-3,6,17-one, was used for all subsequent experiments. Table 2.7 The effect of heating on the yield of 11,17-keto dexamethasone. vial no. treatment ratio 1 1% pyridine, R.T."' 0.66 2 1% pyridine, heat 0.79 3 NaOAc, R.T.* 0.66 4 NaOAc, heat 0.78 I"R.T. = room temperature The effect of heating upon the reaction rate is recorded in Table 2.7. The ratio in the third column is the ratio of the peak area of 11,17-keto dexamethasone to the peak area of the recovery standard (4-androstene-3,6,17-trione) that was added to each oxidation mixture just prior to the clean-up step. Comparison of this ratio for each reaction, rather than the absolute peak areas of 11,17-keto dexamethasone, ought to compensate for losses incurred during transfer steps following the oxidation and for variation in GC injection volume. An increase in the rate of oxidation is observed upon heating both when pyridine is used as the base and when NaOAc is used. The increase in rate is ca. 17%. Due to the boiling point of CH2C12, it is not practical to raise the temperature above 60°C. It should be mentioned that using pyridine causes broadening of the PCC band on 54 the silica column during the clean-up step. Therefore, the use of pyridine could be a disadvantage if the pyridine causes the PCC and the 11,17-keto dexamethasone to co- elute from the silica column. For this reason, and because there appears to be no reaction yield advantage in using pyridine, sodium acetate will be used in all subsequent PCC oxidations. 3. Determination of the Yield of 11,17-Keto Dexamethasone The yield of 11,17-keto dexamethasone from a 5.5 hour oxidation of 50 1.1g of dexamethasone was calculated as 71% (injection standard deviation = 2.4) from the FID standard curve. After 5.5 hours no unoxidized dexamethasone was seen by HPLC or TLC. However, the HPLC chromatogram did contain extra peaks one of which may have been due to a partially oxidized intermediate. (HPLC preparative collection of the peaks followed by direct probe EI-MS was done to determine if any were a partially oxidized form of dexamethasone. However, the presence of contaminants, probably from the concentration of the mobile phase, obscurred the spectra making interpretation difficult. To identify an intermediate oxidation product a larger quantity would be needed.) In order to determine if the reaction yield could be increased beyond the 70% range without a prohibitively long reaction time, another experiment was done in which 167 ng of dexamethasone was oxidized for different lengths of time. The results of this experiment are shown in Table 2.8. 55 Table 2.8 Yield of 11,17-keto dexamethasone by PCC oxidation vs. time. reaction time relative yield 4’50" 74% 8’ 78% 9’ 10" 100% The yield after nine hours was significantly higher than the yield after both 8 and 5 hours. However, since the yield after 5.5 hours was found to be in the 70% range in the earlier experiment, the extra four hours required to increase the yield another 25% were not considered to be worth the extra time. Therefore, a reaction time of 5-6 hours was judged to be sufficient. 4. The Effect of Extracted Plasma Matrix on the Oxidation of Dexamethasone It is possible that the components of the plasma which are extracted along with dexamethasone during the initial isolation step of the method could have an effect on the oxidation of the dexamethasone. Although a good yield may be achieved by a 5-hour PCC oxidation of a pure standard of dexamethasone, the presence of the plasma matrix could decrease the yield considerably. In order to determine the effect of the extracted plasma matrix, triplicate oxidations of 50 ug of dexamethasone were done within the plasma extract. Three milliliters of plasma was extracted using a C18 solid phase extraction column as described in section IIA. The extract was divided into three portions and 50 ug of dexamethasone was added to each portion. For comparison triplicate oxidations of 50 pg of dexamethasone without the plasma matrix were done concurrently. For all of these oxidations NaOAc was used as the base and the reactions were heated to 57°C. After 21/4 hours the recovery standard (4-androstene-3,6,17-trione) was added to each of the vials. The contents of each vial were placed on separate 500 mg 5 6 silica SPE columns. The steroids were recovered with ethyl acetate. Table 2.9 shows the results obtained by GC-FID analysis of these samples. Table 2.9 Ratio of GC-FID responses for 11,17-keto dexamethasone and the internal standard, 4-androstene-3,6,l7-one, for oxidations conducted with and without presence of plasma matrix extract. s= std. deviation of 3 injections. vial no. plasma area ratio mean area ratio 1 yes 0.536 (s=.002) 2 yes 0.59 (s=.01) 0.56 (s=.03) 3 yes 0.554 (s=.006) 4 no 0.614 (s=.002) 5 no 0.664 (s=.004) 0.65 (s=.03) 6 no 0.657 (s=.004) The results of the experiment in which dexamethasone was oxidized in the presence of plasma extract are recorded in Table 2.9. The last column in Table 2.9 contains the means of the two sets of triplicate reactions. These means are significantly different by the t-test at a confidence level of 99%. In other words, there is only a probability of 1% that the difference between the two means is due to random flucuation. This indicates that the production of 11,17—keto dexamethasone is slower in the presence of the plasma matrix. The ratio of the mean ratios is 0.87 which indicates that in the presence of plasma matrix the reaction is only 87% as efficient. The effect of the plasma matrix upon the oxidation of dexamethasone may be even more pronounced when the concentration of the dexamethasone is much less than the concentrations of the matrix components as it would be in real samples. Because of this, reaction times longer than 5 hours may be necessary. This experiment was repeated at the 5 ng/ml level; however, the results by GC-ECD were inconclusive due to interferences from the plasma matrix obscurring the 11,17-dexamethasone and internal standard peak areas. Unfortunately, when this experiment was conducted no mass spectrometer having adequate sensitivity was available to more selectively analyze these samples. 57 5. Oxidation of 6B-Hydroxydexamethasone The compound 6B-Hydroxydexamethasone differs from dexamethasone only by the introduction of the 6B-hydroxy group. It was expected that oxidation of 6B- hydroxydexamethasone would also yield a highly electrophilic product which would have the structure shown in Figure 2.6. Approximately forty micrograms of 6B-hydroxydexamethasone was oxidized in CH2C12 by PCC/celite with NaOAc, at 60°C for about 12 hours. Gas chromatography with ECD yielded a large peak having a retention time very close to that of 11,17-keto dexamethasone. The identity of the peak was confirmed to be 6,11,17—keto dexamethasone by GC-EI/MS. The EI mass spectrum is shown in Figure 2.7a. The ECNI mass spectrum in shown in Figure 2.7b. The base peak, at m/z 324, in the ECNI spectrum represents the loss of HF from the intact molecule. No peak representing the molecular anion is present. Also, the 6,11,17-keto dexamethasone apparently undergoes a reaction under ECNI conditions to produce an ion of mass 326. Reaction of the (M- HF)' with hydrogen radicals could account for the peak at m/z 326. These types of reactions have been reported for aromatic compounds under ECNI conditions (20). 58 Oxidation Figure 2.6. Oxidation of 6B-Hydroxydexamethasone to 6,11,17-keto dexamethasone. 59 145 tom R l 69 . f ‘ 135 ' ‘ 106 4 - a 50‘ 57 115125 17 - t . 81 95 153 . i j 13 185 199210221i V i e 0‘ A 50 too 150 200 _ b I U D n . d A a , n . g 245 259 279 288296 250 300 350 400 m/z ' 324. 4 TxE+o4 100- .—5 89- 326. 4 -4 r__ 60-1 -3 40- __2 20- 1.4 309.3 328.4 1 t— 0 11. L A 1 i m ' 1 m ' 1 280 300 320 340 360 380 Figure 2.7. Mass spectra of oxidized 6B-hydroxydexamethasone obtained by a) electron impact ionization and b) electron capture negative ionization. 6 0 An experiment was conducted to determine the relative yield of 6,11,17-keto dexamethasone vs. time. These results are shown in Table 2.10. For these reactions 194 ng of 6B-hydroxydexamethasone was used. The reaction conditions used were the same as those determined to produce a good yield of 11,17-keto dexamethasone from dexamethasone. The compound 4-Androstene-3,6,17-one was added to the reactions prior to the silica step to serve as an internal standard to correct for transfer losses. Table 2.10 Relative yield of 6,11,17-keto dexamethasone vs time. reaction time rel. %yield 4.5 hr. 86 6.0 hr. 93 7.5 hr. 100 The highest relative yield of 6,11,17-keto dexamethasone was achieved after 7.5 hours. The actual percent yield of 6,11,17-keto dexamethasone could not be determined as no standard of 6,11,17-keto dexamethasone was available. Such a standard could not be synthesized in large quantity as only 2mg of 6B-hydroxydexamethasone was available (a gift from C. Kilts of Duke University). The percent yield for the 7.5 hour oxidation may be estimated to be on the order of 15% by assuming that the ECD responses of 6,11,7-keto dexamethasone and 11,17-keto dexamethasone are approximately 3 to 1. This assumption is based on the observation (21) that the relative ECNI responses of 1,4- androstadiene-3,6,11,17-one and 1,4-androstadiene-3,11,17-one are 3 to 1. The only difference between these molecules and the oxidation products of 6B- hydroxydexamethasone and dexamethasone is that the dexamethasones have 9a—F and l6oc-methyl groups. No other major peaks were seen in the GC-ECD chromatogram. No peak was seen at the retention time corrresponding to the 17-keto analog of 6B- hydroxydexamethasone (a standard was prepared by NaBiO3 oxidation of 61 6B-hydroxydexmethasone). If unoxidized 6B-hydroxydexamethasone were present in the samples it would be expected that the 17-keto analog would be observed in the chromatogram as a result of thermal decomposition of 6B-hydroxydexamethasone in the GC injection port. C. Removal of Excess Reagent and Isolation of Oxidized Dexamethasone and 613- Hydroxydexamethasone It is important to remove the oxidant completely when isolating the reaction products so that the solution of products will not degrade over time. Complete removal of oxidant also prevents damage to the GC column. This section discusses methods for PCC removal and optimal recovery of 11,17-keto dexamethasone and 6,11,17-keto dexamethasone. 1. Silica vs. Solvent Extraction An experiment was done to compare the efficiency of the removal of PCC and the recovery of 11,17-keto dexamethasone by ethyl acetate solvent extraction with the efficiency by silica solid phase extraction, employing ethyl acetate as the eluent. The recoveries of 11,17-keto dexamethasone obtained by the two methods were not significantly different (both were ca. 95%). Because solvent extraction involves at least three extraction steps and emulsions can form at each step, it is a more time consuming method than solid phase extraction. Therefore, solvent extraction was not used. 2. Comparison of Sephadex LH-20 and Silica Sephadex LH-20 had been used previously in this laboratory for removal of the excess pyrzCr03 and isolation of oxidized dexamethasone. Removal of this oxidation 62 reagent was efficient using this procedure, but the recovery of oxidized dexamethasone was never assessed. Sephadex LH-20 was found to effectively remove PCC as well. In order to determine the most efficient method for the recovery of 11,17-keto dexamethasone, the efficiency of recovery using Sephadex LH-20 was compared to the efficiency achieved using silica columns. The results of the comparison of Sephadex LH-20 and silica are shown in Table 2.11. Table 2.11 Recovery of 11,17 -keto dexamethasone by Sephadex LH-20 and by silica SPE column. column type eluent % recovegl Sephadex 3.5 ml ethyl acetate 96% silica 3.5 ml ethyl acetate 104% silica 3.5 ml ethyl acetate-CH2C12 (3:1) 80% The results in Table 2.11 indicate that both silica and Sephadex give quantitative recovery of 11,17-keto dexamethasone. The recovery of 11,17-keto dexarrrethasone is somewhat less if an equal volume of a less polar eluent is used, as seen in the third entry of Table 2.11. Since the silica SPE columns are commercially made, the reproducibility of the recovery ought to be better with these columns than with the home-made Sephadex LH- 20 columns. Because of this and because the use of Sephadex LH-20 has no advantage for recovery of 11,17-keto dexamethasone, silica SPE columns were used for all subsequent experiments. 3. Optimization of silica solid phase extraction of 11,17-keto dexamethasone and 6,1 1,17-keto dexamethasone Although much selectivity for the detection of 11,17-keto dexamethasone and 6,11,17-keto dexamethasone is provided by the chemical oxidation and the 63 GC-ECNI/MS (SIM) analysis, the silica step may be used to provide another dimension of selectivity. Isolation of 11,17-keto dexamethasone and 6,11,17-keto dexamethasone from other potentially interfering components in oxidized plasma matrix may allow a less selective means of introducing the sample into the mass spectrometer, such as the direct inlet probe (DIP). The use of DIP would simplify and shorten the mass spectrometric analysis. In order to isolate 11,17-keto dexamethasone and 6,11,17-keto dexamethasone more selectively, the volume and the polarity of the eluent was varied and fractions were collected. The following experiment was done to determine the volume of ethyl acetate necessary to elute all of the 11,17-keto dex from the silica column. After a 5 1/2 hour PCC oxidation of 50 ug of dexamethasone, the reaction mixture was transferred to a 500 mg silica Bond-Elut column. Fractions of the eluate were collected and analysed by normal phase HPLC to determine the presence of product. These results are shown in Table 2.12. Table 2.12 HPLC analysis of fractions from silica column used for removal of PCC following oxidation of dexamethasone. fraction eluent 11.17-keto dex 1 1 ml CH2C12 0% 2 2.5 ml ethyl acetate 97.5% 3 2.5 ml ethyl acetate 2.5% 4 80% MeOH in ethyl acetate 0% The HPLC analysis of the silica fractions shown in Table 2.12 indicate that nearly all the 11,17-dex is eluted from the 500 mg silica column with 2.5 ml of ethyl acetate. Previous to this experiment, 4-8 ml of ethyl acetate had been used to recover 11,17 -keto dexamethasone from reaction mixtures. The use of a smaller volume of ethyl acetate will eliminate the elution of more polar matrix components from the column. 64 The use of an eluent less polar than ethyl acetate was investigated for the recovery of both 11,17-keto dexamethasone and 6,11,17-keto dexamethasone. The use of a CH2C12 wash was implemented, as well, to elute matrix components of lower polarity than 11,17-keto dexamethasone and 6,11,17-keto dexamethasone. These results are shown in Table 2.13. For this experiment 3 ml of plasma were spiked with 105 ng of dexamethasone and 78 ng of 6B-hydroxydexamethasone. The dexamethasone and 6B- hydroxydexamethasone were extracted from the plasma and oxidized with PCC according to the procedures described previously in this chapter. After the oxidation the reaction mixture was divided into three approximately equal portions. Each portion was placed on a separate silica column. Different eluents and volumes of eluents were applied to each column. Only the results from analysis of fractions collected from one of the columns are shown in Table 2.13. Due to the low levels of dexamethasone and 6B- hydroxydexamethasone added to the plasma, mass spectrometry was used to provide more selectivity for the analysis of the fractions than is provided by the GC-ECD. Table 2.13 DIP-ECNI/MS analysis of silica SPE column for recovery of 11,17-keto dexamethasone and 6,11,17-keto dexamethasone. fraction eluent 11.1flceto dex 6,11,17-keto dex 1 2 ml CH2C12 0% 0% 2 0.7 ml 5% acetone-CH2C12 0% 0% 3 2.8 ml 5% acetone-CH2C12 100% 100% The results in Table 2.13 indicate that the CH2C12 wash does not elute 11,17-keto dexamethasone or 6,11,17-keto dexamethasone and that 5% acetone-CH2C12 is polar enough to elute both of these compounds. The percentages in Table 2.13 represent the portion eluted in the fraction relative to the total amount eluted in all three fractions. It is possible that some of the 11,17-keto dexamethasone and 6,11,17-keto dexamethasone still remained on the column after the collection of the third fraction. 65 Another experiment was done to determine the volume of 5% acetone-CH2C12 necessary to recover all of the oxidized 6B-hydroxydexamethasone. Since 6,11,17-keto dexamethasone is more polar than 11,17-keto dexamethasone, the volume necessary to recover 6,11,17-keto dexamethasone will most likely be adequate to recover all of the 11,17-keto dexamethasone as well. Approximately 40 mg of 6B-hydroxydexamethasone was oxidized with PCC. The results of the GC-ECD analysis of the silica fractions is shown in Table 2.14. Table 2.14 GC-ECD analysis of silica fractions for isolation of 6,11,17-keto dexamethasone. fraction eluent 6,11,17-keto dex 1 1.5 ml CH2C12 4% 2 4 ml 5% acetone-CH2C12 89% 3 4 ml ethyl acetate 7% A volume of 4 ml of 5% acetone-CH2C12 recovered 89% of the 6,11,17-keto dexamethasone. A small percentage was eluted by the CH2C12 wash. In the previous experiment, none of the 6,11,17-keto dexamethasone eluted in the 2 ml wash. The most likely reason for this discrepancy is that a much greater amount of the sample (ca. 40 mg) was applied to the column in the second experiment than in the frrst experiment (ca. 26 ng). The greater amount of sample would have a much broader elution band than the small amount of sample. Four milliliters of 5% acetone-CHzClz were used to elute 11,17-keto dexamethasone and 6,11,17-keto dexamethasone from silica. This ensures that nearly all the 11,17-keto dexamethasone and 6,11,17-keto dexamethasone are eluted from the silica column, and minimizes the elution of components more polar than these analytes. Only one and one-half milliliters of CH2C12 were used to minimize the possible elution of 11,17-keto dexamethasone and 6,11,17-keto dexamethasone in the CH2C12 wash. 66 III. Conclusion Conditions for the isolation and oxidation of dexamethasone have been studied extensively to establish optimal recovery of 11,17-keto dexamethasone. Some preliminary work for the isolation and oxidation of 6B-hydroxydexamethasone has also been completed. Extraction of both dexamethasone and 6B-hydroxydexamethasone from plasma was most efficient using the C18 solid phase extraction method of Kilts and Ritchie modified by the implementation of a 15% methanol-water wash, rather than the 38% methanol-water wash. Recovery of dexamethasone and 6B-hydroxydexamethasone (at the 200-400 11 g level) was approximately complete using the modified method. The best conditions of those investigated for the conversion of dexamethasone to 11,17-keto dexamethasone were oxidation with PCC ground up with celite, with NaOAc buffer, in CH2C12, stirred at 60°C. A yield of 70% 11,17-keto dexamethasone could be achieved after 5.5 hours using these conditions. Higher yields could be obtained at longer reaction times. The yield of 6,11,17-keto dexamethasone from the PCC oxidation of 6B-hydroxydexamethasone was not optimized, however, 6,11,7-keto dexamethasone was produced under the conditions used for the oxidation of dexamethasone. Recovery of 11,17-keto dexamethasone and 6,11,17-keto dexamethasone from the reaction mixture may be accomplished through the use of a silica solid phase extraction column. Complete recovery can be achieved using ethyl acetate as the eluent. To achieve a nearly quantitative recovery of 11,17-keto dexamethasone and 6,11,17-keto dexamethasone, while avoiding co-elution of some of the matrix components, a 1.5 ml CH2C12 wash followed by collection of 4 ml of 5% acetone-CH2C12 is recommended. 10. ll. 12. 13. 14. 15. l6. l7. 18. REFERENCES B. J. Carroll, M. Feinberg, J. F. Greden, J. Tarika, A. A. Arbala, R. F. Haskett, N. M. James, Z. Kronfol, N Lohr, M. Steiner, J. P. de Bigne, and E. Young, Arch. Gen. Psychiatry, 38, 5-22, 1981. M. T. Lowy and H. Y. Meltzer, Biol. Psychiatry, 22, 373-385, 1987. G. F. Johnson, G. Hunt, K. Kerr, and I. Caterson, Psychiatry Res, 13, 305-313, 1984. G. W. Arana, R. J. Workman, and R. J. Baldessarini, Am. J. Psychiatry, 141, 1619-1621, 1984. B. S. Cham, B. Sadowski, J. W. O’Hagon, C. N. de Wytt, F. Bochner, and M. J. Eadie, Ther. Drug Monitoring, 2, 373-377, 1980. Y. Kasuya, J. R. Althaus, J. P. Freeman, R. K. Mitchum, and J. P. Skelly, J. Pharm. Sci, 73, 446-451, 1984. K. Minagawa, Y. Kasuya, S. Baba, G. Knapp, and J. P. Skelly, J. Chromatogr, 343, 231-237, 1985. Minagawa, Y. Kasuya, S. Baba, G. Knapp, and J. P. Skelly, Steroids, 47, 175- 188, 1986. D. S. Skrabalak, K. S. Cuddy, and J. D. Henion, J. Chromatogr, 341 , 261-269, 1985. personal communication with Clinton Kilts of Duke University Medical Center, Departments of Pharmacology and Psychiatry, Durham, NC. Guor Ron g Her, Mass Spectrometric Studies of Molecules Modified for Analysis by Electron Capture Negative Chemical Ionization, Ph.D. Dissertation, Michigan State University, Department of Chemistry, 1985. K. S. A. Razzak and K. A. Hamid, Biomed. Mass Spectrom, 7, 505-510, 1980. C. H. L. Shackleton and J. 0. Whitney, Clin. Chim. Acta, 107, 231-243, 1980. P. M. Plezia and P. L. Berens, Clin. Chem, 31, 1870-1872, 1985. F. Freeman, "Oxidation by Oxochromium (VI) Compounds" in Organic Syntheses By Oxidation With Metal Compounds, M. J. Mijs, and C. R. H. I. de Jonge, eds, p.44, Plenum Press, New York, NY, 1986. E. J. Corey and J. W. Suggs, Tetrahedron Lett, 2647-2650, 1975. J .C. Collins, W. W. Hess, and F. J. Frank, Tetrahedron Lett, 3363-3366, 1968. J. M. J. Frechet, P. Darling, and M. J. Farrall, J. Org. Chem, 46,1728-1730, 1981. 67 19. 20. 21. 6 8 H. H. Willard, L. L. Merritt, J. A. Dean, and F. A. Settle, Instrumental Methods of Analysis, Sixth edition, Wadsworth Publishing Company, Belmont, CA, pp. 468- 470, 1981. P. S. Callery, W. A. Garland, and E. K. Fukuda, Org. Mass Spectrom, 24, 385- 390, 1989. G. R. Her and J. T. Watson, Biomed. Environ. Mass Spectrom, 13, 57-63, 1986. CHAPTER 3. COMPARISON OF GC-MS AND MS/MS TECHNIQUES IN THE ANALYSIS OF PLASMA FOR DEXAMETHASONE I. Introduction Described in this chapter is the adjustment of ECNI-MS conditions and MS/MS conditions for the detection of oxidized dexamethasone. Concentrations of dexamethasone in plasma samples from patients undergoing the dexamethasone suppression test (DST) have been determined by these mass spectrometric techniques using the sample preparation methodology described in Chapter 2. These results are compared with those obtained by GC-EI/MS and RIA methods in other laboratories, where aliquots of the same plasma samples were analyzed. The GC-EI/MS method for the determination of dexamethasone is based on the formation of the TMS-enol-TMS ether derivative (1). This method represents the best conventional GC-MS method available for the determination of dexamethasone (2). The structure of this derivative and selected ion current profiles of the (M-CHzOTMS)+ ion at m/z 577 and the molecular ion at m/z 680 are shown in Figure 3.1. The peak at m/z 577 represents the most intense high mass fragment ion in the EI mass spectrum of this derivative. The sensitivity of an assay based on this derivative is limited by the low intensity of the peaks representing these ions. Figures 3.1 and 3.2 compare the sensitivity available through this GC-EI/MS method with the sensitivity available through oxidation and GC-ECNI/MS analysis. The data in Figure 3.1 were obtained from 2 ml of blank plasma that was spiked with 5 ng of dexamethasone. The plasma was extracted using a C18 solid-phase extraction column. The extract was divided into two equal portions. One portion was derivatized to make the TMS-enol-TMS ether and the other portion was oxidized to produce 11,17-keto dexamethasone. The GC-EI/MS selected ion current profiles in Figure 3.1 result from a 69 7 0 6% injection of the derivatized portion; this represents at most 150 pg of the dexamethasone added to the blank plasma. Figure 3.2 shows the selected ion current profile obtained by GC-ECNI/MS analysis of a 5% aliquot of the oxidized portion of the extracted plasma; this represents no more than 125 pg of the original dexamethasone. By comparison of the s/b ratios of the selected ion current profiles by the two methods it is apparent that greater sensitivity ought to be available using oxidation with GC-ECNI/MS analysis. II. Sample Preparation andAnalysis of Plasma for Dexamethasone Using SIM A. Experimental This section describes the experimental details for the analysis of plasma for dexamethasone. Included here is the protocol for sample preparation, as well as the gas chromatographic conditions and the selected ion monitoring parameters that were used. The adjustment of the ECNI parameters is also discussed. 1. Materials and Sample Preparations Chemicals used for the oxidation procedure include pyridinium chlorochromate (Aldrich Chemical Company, Inc.) and anhydrous sodium acetate (J. T. Baker Chemical Company). All solvents used were either analytical or HPLC grade. They were purchased from Burdick and Jackson Division of Baxter Healthcare Corporation, J. T. Baker Chemical Company, EM Science, and Mallinckrodt, Inc. CeliteTM analytical filter aid was purchased from Johns-Manville Products Corporation. Analytical reagent grade urea was purchased from Mallinckrodt, Inc. Solid-phase extraction columns (200 mg C18 PrepSepTM) were purchased from Fisher Scientific. Silica, solid-phase extraction 71 CHOTMS TMSO — _ CH3 0 El 1‘1 MMNJWMWWMMQW WWMML MW m/z 680 J. x l - WI‘WWNIW‘VWR'MWR‘MtWKWW‘M.W#~ 1'4 is is 1'1 18 Time (minutes) Figure 3.1 Selected ion current profiles for detection of the TMS-enol-TMS derivative of dexamethasone within plasma mauix, by GC-EI/MS. 72 EC-Nl - m/z 330 plus m/z 310 Time (minutes) Figure 3.2 Selected ion current profile for detection of oxidized dexamethasone in plasma matrix, by GC-ECNI/MS. 7 3 columns (500 mg) were purchased from Burdick and Jackson Division of Baxter Healthcare Corporation. Dexamethasone was purchased from Steraloids, Inc. l3C6-2H3 dexamethasone was a gift from Clinton Kilts of Duke University. All materials were used without further purification. Silanized glassware was used throughout the procedures. The preferred conditions for sample isolation and oxidation'were established by the studies described in Chapter 2. This sample preparation consists of three steps: plasma extraction, chemical oxidation, and removal of excess oxidation. Before extraction, the internal standard (35 ng of 13C5-2H3 dexamethasone in 50 u] of methanol) was added to each sample. Samples consisted of 1.5 ml of plasma taken from patients at various times following administration of 1 mg of dexamethasone. PrepSepTM solid phase extraction columns containing 200 mg of C18 packing were pre- conditioned with 5 ml of an 8 M urea solution to deactivate any remaining active sites, followed by washes with 5 m1 of methanol and 10 m1 of distilled water. The plasma samples were then applied to the columns followed by washes with 4 ml distilled water and 4 ml of 15% methanol in water. Dexamethasone was eluted with 4 ml of 70% methanol in hexane. All solvents were pulled through the solid-phase extraction columns at flow rates of approximately 1-2 drops/sec through the use of a vacuum manifold, except the final elution which was done by gravity flow to avoid contamination from the vacuum manifold. Following removal of the eluting solvent under nitrogen, the dried residues were each reconstituted in 1 ml of CH2C12 while stirring with a magnetic stirrer to promote dissolution of the extracted material. Approximately ten milligrams of anhydrous sodium acetate were added to the reconstituted extracts. Sodium acetate is added to buffer acid liberated during the oxidation. An excess of pyridinium chlorochromate, finely ground with celite (ca. 1:3 w/w) by using a mortar and pestle, was then added to the solutions. The reaction mixtures were stirred and heated to 50-60‘C for 6 1/2 hours. 74 After completion of the oxidation, the reaction mixtures were directly transferred to 500 mg silica solid phase extraction columns that had been prewashed with several column volumes of CH2C12. The columns were then washed with 1.5 ml of CH2C12. The oxidation product of dexamethasone was eluted with 4 ml of 5% acetone in CH2C12. After removal of the 5% acetone-CH2C12 under nitrogen, the residue was reconstituted in 100 [.11 of ethyl acetate. The GC-ECNI/MS analysis required 1 to 3 pl of sample. For preparation of the calibration curve, 13C5-2H3-dexamethasone (35 ng) was added to each of several 1 ml aliquots of blank plasma obtained from the Red Cross blood bank. Following this, additions of 0.0, 0.209, 0.418, 1.67, 4.18, 8.36, and 16.7 ng of dexamethasone, in 25 to 100 pl of methanol, were made to the 1 ml aliquots of plasma. . Each standard was prepared in duplicate. The extraction, oxidation, isolation, and analysis procedures for these standards were identical to the procedures used for the clinical plasma samples. Another standard was made to be used in the assessment of the precision and accuracy of the methodology. For this standard, one milliliter of blood bank plasma was spiked with 35 ng of 13C6-2H3-dexamethasone, followed by addition of 2.09 ng of dexamethasone. This standard was prepared in duplicate and analysed in triplicate by GC-ECNI/MS with SIM. 2. Adjustment of Instrumental Parameters All GC-ECNI/MS data were acquired with a Finnigan TSQ70 directly coupled with a Varian GC. For selected ion monitoring, windows of 0.8 u about the ion currents at m/z 339.2, 319.2, 330.2 and 310.2 were monitored for a dwell time of 50 milliseconds each. Ion current at m/z 324.2 was also monitored to detect the oxidation product of 75 6B-hydroxydexamethasone. This 0.8 u window was chosen to maximize the detection of the ion current at these masses without detection of ion current at adjacent masses. Ammonia (Matheson, 99.99% purity) was used as the reagent gas at a source pressure that corresponded to a foreline pressure of 9300-9500 mtorr. This pressure was obtained from the convectron gauge that is located on the foreline before ion source and the turbomolecular pumps. The pressure reading at this point is considerably higher than the actual pressure inside the ion source chemical ionization volume. There was no available means by which to measure the actual ion source pressure. The operator’s manual for the TSQ70 recommends a pressure of about 0.5 torr be used for CI conditions. Because good sensitivity was obtained at convectron readings of 9000-9900 mtorr, it was assumed that this foreline pressure corresponds to an ion source pressure on the order of 0.5 torr. Both CH4 and NH3 were investigated as the modifying gas. The sensitivity for . the detection of 11,17-keto dexamethasone appeared to be same regardless of the gas that was used; however, when CH4 was used higher background was experienced. Intense peaks at m/z 233 and m/z 235 in a ratio of 1:3 were observed, particularly when CH4 was used. The masses of these ions and the isotope pattern correspond to ReO3-. Others have noted a high background due to these isotopes under CH4 ECNI conditions (3). The source of the rhenium is the filament used to generate the primary electron beam. The ion source pressure was roughly optimized by comparing the ion current resulting from 3.6 ng of 11,17-keto dexamethasone obtained at three different pressures of NH3. The pressures investigated corresponded to readings of 9000, 9500, and 9900 mtorr on the convectron gauge. The intermediate pressure corresponding to a reading of 9500 mtorr on the convectron gauge resulted in the most intense signal. Therefore, each time the instrument was used, the pressure was adjusted to operate in this region. No rigorous attempt was made to optimize the pressure because the sensitivity of the instrument seemed to be adequate at each of these three pressures. The ion source temperature used for the analyses was 120°C. As the ion source 7 6 temperature increases, the relative abundance of the fragments increase and the abundance of the molecular ion decreases until none is present. As the temperature of the ion source rises, the total ion current increases also. The source temperature of 120°C was chosen in order to balance the tradeoff between the absolute abundance of the ion current and the abundance of the molecular anion. At this temperature the relative abundance of the molecular ion was about 10% of the (M-HF)' fragment ion. More selectivity could be provided by monitoring both the molecular anion and this major fragment ion. The ionization voltage used was 100 eV. The ion current resulting from 3.6 ng of 11,17-keto dexamethasone was compared at ionization voltages of 70, 100, 120, and 150 eV. The greatest ion current was obtained at 100 eV, although the differences in ion current at the different ionization voltages were not very significant. The difference in ion current obtained upon changing the ionization voltage seemed to be more pronounced at the lower ion source temperature (100°C) than at the higher ion source temperature (140°C). The ionization current used for all analyses was 300 uA. The electron multiplier voltage was 1.2 kV. The direct capillary inlet was heated to 280°C. Splitless injection was used to introduce samples. The injector temperature was 280°C. A 30 m x 0.25 mm x 0.25 pm fihn thickness DB-S fused silica capillary column was used for all analyses. The temperature program employed was 60°C-260°C at 40°C/min, then to 280°C at 4°C/min. After each analysis, the oven was ramped up to 300°C for 3-5 minutes to bake out any remaining sample components. The GC column head pressure was 10 psi g. 77 B. Results and Discussion The electron capture negative ion mass spectrum of 11,17-keto dexamethasone is shown in Figure 3.3. The peak at m/z 330 represents the molecular anion, the peak at m/z 310 represents the neutral loss of HF from the molecular species, and the peak at m/z 295 represents the loss of a methyl radical from the (M-HF)- ion. The chemical oxidation product of dexamethasone, 11,17-keto dexamethasone, is shown in the inset in Figure 3.3. Both the 16a- and 16B-methyl epimers of 11,17-keto dexamethasone are observed. About 20% epirnerization occurs after formation of the 17- keto group due to the acidity of the 16B-hydrogen alpha to the 17-ketone. The presence of pyridine (from PCC), and possibly the sodium acetate, promote enolization of the ketone, leading to epimerization. Without the addition of sodium acetate, however, the extent of epirnerization is about the same. These epimers can be separated by capillary gas chromatography. The presence of the two peaks (a well resolved doublet) can be seen as an advantage because the pattern is easily recognizable in the presence of other chromatographic peaks. The limit of detection of a pure standard of 11,17-keto dexamethasone was about 160 fg with s/n of 7 (n = noise calculated at the top of the peak) by selected ion monitoring. The ion current profile at m/z 310 is illustrated in Figure 3.4. A blank injection of solvent prior to the injection of 160 fg of 11,17-keto dexamethasone was monitored to ensure that the signal observed for 160 fg of 11,17-keto dexamethasone was not the result of residual 11,17-keto dexamethasone being "washed off" the injector or septum. The detection limit is improved by about 50 times over that reported previously by Her and Watson (4) with the HP5985 modified in-house for negative ion detection. Also contributing to the improved sensitivity observed with the TSQ70 is the fact that the GC column is directly interfaced. The HPS985 had an open split GC-MS interface which decreased the amount of analyte entering the ion source. .ocommfioamxow 982%.: .«o 8309.5 93 25350838 98.-S.S «o 883on 9.38 20m m.m 2:me NE 3m 98 an . o 5 8m 8m 0 .._.. pb...r...-..p....._p. ..._....__...»_._..... _ [mom 0 mam low m -A IQHEE 18 r8 .2 0mm Ex: 18 roe o8 79 ‘00 m/z 310 s/n i 7 80 ' I 1 40" 20- TT‘IIUUUITU'Ith'IIIVUUI'TTT‘U'IIIUUUIIII 11 11:30 12 12:30 time Figure 3.4 GC-ECNI/MS selected ion current profile from 160 fg of 11,17- keto dexamethasone standard. 8 O The calibration curve, shown in Figure 3.5a, was obtained by plotting the ratio of the peak height at m/z 310 to that at m/z 319 ((M-HF)- from the M+9 internal standard) for the 16-0t epimer of 11,17-keto dexamethasone against ng of dexamethasone added to the calibration curve standards. Peak heights rather than peak areas were used to construct the curve because of occasional difficulty in defining the baseline due to co- eluting interfering substances that produced ion current at either m/z 319 or m/z 310. Linear regression of the data provided the equation to be used for quantification of dexamethasone in real plasma samples (y = 0.0259 + 0.0360x). The correlation coefficient for the curve was 0.998. The plot shown in Figure 3.5a confirms the linear relationship of the data throughout this range of dexamethasone concentrations. The high y-intercept value is partly due to the presence of unlabelled dexamethasone in the labelled internal standard, 13C6-2H3 dexamethasone. This internal standard was received as a gift and, unfortunately, it was contaminated with unlabelled dexamethasone upon arrival. Although the reagents available for synthesis of 13C- labelled compounds are less than 100% enriched, the source of the unlabelled dexamethasone in the 13C5-2H3 dexamethasone was probably not due to the synthetic process. If this were so, molecular ions at m/z 331-338 would be observed. Molecular ions at these masses are not observed, except for naturally occurring isotopic peaks at m/z 331 and m/z 332. The limit of detection for dexamethasone (LOD) in plasma may be calculated from the y intercept of the calibration curve and the standard deviation of the y intercept. A generally accepted criterion for the limit of detection is YIod = yb + 381) (5), where yb is the value of the y intercept of the calibration curve or the calibration curve blank, sb is the standard deviation associated with the measurement of the blank, and YIod is the lowest y value (ratio of m/z 310 to m/z 319 in this case) that will reliably represent the presence of the analyte. Duplicate SIM analyses of duplicate blanks (plasma samples containing only internal standard), gave a standard deviation of 0.003 (%RSD = 11). 81 .o .o O\ 00 I (m/z 310) /I (m/z 319) o .b 20 0.2 0.0 d ‘ 0 10 ng of dexamethasone _ b) 0.3 a y = 7.55 e-2 + 0.175 x R42 = 0.997 F; E 0.2 ’9? m 0.1 N E 0.0 I? v I v I r I V I 0.0 0.2 0.4 0.6 0.8 ng of dexamethasone Figure 3.5 Calibration curves for determination of dexamethasone in plasma b GC-ECNI/MS with SIM using a) 35 ng, and b) 3.5 ng of 1 C6-2H3-dexamethasone internal standard. 1.0 82 From the y intercept of the calibration curve and this standard deviation, the ylod is calculated as 0.0350. This y value corresponds to an x value of 0.25 ng dexamethasone/35 n g internal standard. This detection limit will be adequate to detect dexamethasone reliably in all but the least concentrated plasma samples, since the expected concentration range is from 0.1 to 4 ng/ml. This detection limit does not approach the actual detection limit of the instrument (160 fg pure standard gives s/n = 7). Rather, the limitation comes from the poor precision experienced in measurement of the peak height ratio of the blank, and the low slope of the calibration curve. A rather large excess of internal standardwas added in order to serve as a carrier as well as an internal standard. The sensitivity of the assay might be increased if a structurally-similar compound was used as the internal standard or a smaller amount of the labelled internal standard was used. The smaller amount of internal standard would increase the slope of the calibration curve, leading to a lower limit of detection and higher sensitivity. This would allow more reliable detection of the lowest levels of dexamethasone in the plasma samples. To demonstrate this, another set of standards containing 0.0, 0.209, 0.418, and 0.836 ng of dexamethasone was processed. The procedures used were identical to those described above, except that only one-tenth the amount of internal standard (3.5 ng) was added. No other compound was added to act as a carrier. The calibration curve obtained from the GC-ECNI-SIM analyses of these standards is shown in Figure 3.5b. The equation obtained upon linear regression of the data is inset in Figure 3.5b. The standard deviation associated with the measurement of the peak height ratio of the blank was 0.009 (12% RSD). The limit of detection was calculated as 0.15 ng/ml from the equation for the ylod that was stated above and the y-intercept of the calibration curve shown in Figure 3.5b. This represents a 40% decrease in the LOD. Another way to decrease the LOD would be to use a structural analog for the internal standard, such as flumethasone. Flumethasone is identical to dexamethasone except that this drug contains an additional fluorine at C6. The use of a structurally 8 3 similar internal standard ought to result in a much lower blank value for the calibration curve. When a structurally similar internal standard is used, the standard deviation of the blank will be lower because the blank value is lower. This will be true as long as the %RSD for the measurement of the blank is not higher when using a structural analog as the internal standard. If the slopes of the calibration curves obtained using the labelled internal standard and the structurally similar internal standard are equal, the decrease in the LOD will be equal to the decrease in the standard deviation of the blank. And, the decrease in the standard deviation of the blank ought to parallel the decrease in the value of the blank, or y-intercept. Unfortunately, for determinations of dexamethasone by oxidation and GC-ECNI- MS, fiumethasone would not be a good choice for the internal standard. The ECNI mass spectrum of the oxidized 11,17-keto fiumethasone contains a peak corresponding to m/z 310 from the (M-2HF)- ion. Because the oxidized products of dexamethasone and flumethasone cannot be completely separated by the GC capillary column, the response at m/z 310 from the internal standard would overlap with the response at m/z 310 from 11,17-keto dexamethasone. This would result in a high blank value. Thus, no advantage would be realized by using flumethasone as the internal standard. If the molecular anions were used instead of the fragment ions for the determination it would be possible to use flumethasone. However, this would decrease the sensitivity of the analysis as the abundances of the molecular anions are lower than the abundances of the fragment ions. For future analyses it is advisable to use the smaller amount of isotopically- labelled internal standard. However, it was not possible to repeat the analyses of these DST plasma samples as all of the sample was consumed. The accuracy of the assay for dexamethasone in plasma was determined by triplicate analyses of duplicate samples "spiked" to a concentration of 2.09 ng/rnl. For the determination of dexamethasone in real plasma samples, these data were included in the calibration curve. 84 Table 3.1 Determination of dexamethasone in 1 ml blank plasma "spik " with 2.09 ng of dexamethasone by GC-ECNI-MS with SIM. sam le trial 1 trial 2 trial 3 mean s %RSD %error 7A 2.01 2.21 2.05 2.1 0.1 5.1 0.0 7B 1.97 1.79 1.72 1.8 0.1 7.1 12.4 overall 2.0 0.2 9.1 6.2 Table 3.1 indicates that the accuracy of the assay is reasonably good at the 2 ng/ml level. The mean of all six determinations was within 6.2% of true value. The precision of the instrumental method is 5 to 7 %RSD. The precision of the overall procedure is on the order of 10 %RSD. The variation between 7A and 7B is also cause by the variation experienced in the additions of the dexamethasone and the 13C5-2H3 dexamthasone to the two aliquots of blank plasma. Results from the SIM analysis of a real plasma sample (P5) are shown in Figure 3.6. Selected ion current profiles for the M- and (M-I-IF)- ions produced from dexamethasone and the internal standard are shown. Substances present in the plasma matrix produced very little ion current that interfered with the detection of 11,17-keto dexamethasone. The peak height ratio of m/z 310 to rn/z 319 of the first epimer was used to calculate the dexamethasone concentrations of the ten DST plasma samples that were analyzed. The concentrations are shown in Table 3.2. The peaks at m/z 330 and m/z 339, representing the molecular anions, were also monitored to provide more specificity for the analyses. The selected ion current profile at m/z 324 for detection of oxidized 6B- hydroxydexamethasone in plasma sample P5 is shown in Figure 3.7 along with the profiles for m/z 310 and m/z 319. The arrows indicate the retention time of the 160t- and l6B-methyl epimers of 6,11,17-keto dexamethasone. Because no calibration curve for 6B—hydroxydexamethasone was available, the concentration of 6B-hydroxy- dexamethasone was not determined. However, these data show that it is possible to 85 Selected ion current profiles m/231O : m/2330 .é‘ - (D C .1 2 .5 ' 9 f; ‘ m/z 319 1306:2113 internal standard ' 72 - - . (M-HF) - _ A _ A _ 1 m/2339 M' *1!“r'rfTrTrrfT‘:r'rrr'r'r‘I‘r‘rrjfrfirfiT 9 10 11 12 13 14 15 time Figure 3.6 Selected ion current profiles from GC-ECNI/MS analysis of plasma sample P5 from patient undergoing the DST. 86 .mm 0383. «Emma 5 53:89.8 .oaomanmeoaonEéo Ea .ocomafiofiaxoe-mm~bvm_ 658525er 3528 .«o 2.2 -Eméé 05 wcucomoaou .vmm 23.3 .on NE a $an 2:88 :2 8823 $628.00 00mm ommm oomm omkm 80km omen 000m 3 28$ onwm Gawm l—rbrrPPbbLbP—bupb-PblbbFPnbb_bbhlP—thLlFP~blF~® :35. mo+mx eNmu~\e Gamun\a ION 16+. r80 row roaa X 87 detect 6B-hydroxydexamethasone in human plasma through oxidation and GC—ECNI-MS analysis. Table 3.2 Comprison of concentrations of dexamethasone in n g/ml of plasma as determined by GC-ECNI-MS with SIM and by conventional GC-EI/MS of the TMS-enol-TMS derivative. Sample ECNI-SIMa EI-SIMb 31AM (n=9) P2 0.63(s=0.07, n=2) 0.83 0.66(s=0.01) P5 0.62(s=0.09, n=3) 0.30 0.4 (s=0.1) P11 3.2 (3:04, n=2) 0.50 0.9 (s=0.3) P18 30 (S: 4, n=2) 28.01 21 (5:8) P3 0.18(s=0.01, n=4) 0.24 0.3 (s=0.1) P4 0.06(s=0.04, n=2) 0.47 0.2 (s=0.1) P16 0.10(s=0.12, n=2) 0.15(LOD) 0.10(s=0.04) P14 0.04(s=0.01, n=2) 0.15(LOD) 0.12(s=0.04) P15 0.06(s=0.06, n=2) >.L >>>.-.p.- .pr....>> >.-—»>P»> .p.>»r.>l«_>>..~» O mam 0mm row Wow L2 0mm 18 his 12: 05 95 parent ion of oxidized dexamethasone. During CAD, only a few electron volts of energy are required to induce fragmentation to the (M-HF)' species. This suggests that the (M- I-IF)' species is more stable than the molecular parent ion. Multiple collision conditions, generated with higher collision gas pressure is favorable, presumably to ensure that all the parent ions do undergo collision with argon. A plot of the absolute abundance of m/z 310 at each energy and pressure that was investigated is shown in Figure 3.9. From Figure 3.9, it is evident that the optimum detectability of the daughter fragment occurs at a collision gas pressure of 1.4 mtorr argon and a collision energy of 3 eV. In order to determine the extent to which the collision gas present in Q2 attenuated the total ion beam reaching the detector, the abundance of the total ion current resulting from MS/MS of m/z 330 was measured with and without collision gas in Q2. This was repeated at a series of collision energies, between 2 and 20 eV. The peak areas . of the chromatographic profiles were measured and compared to determine the collection efficiency at each energy, with and without collision gas. The collection efficiency is defined as the percentage of the ion current transmitted, either as unreacted parent ion or as daughter ions, through the second quadrupole collision cell. The ratio of the ion current detected at a given collision energy with and without gas in Q2 is the collection efficiency. At collision energies less than 5 eV, the collection efficiency was near 100%, however at higher energies it was diminished. This was true for the three pressures that were examined. This decreased efficiency may be attributed to poorer transmission properties of the rf-only second quadrupole when operated at high collision energy since the scattering angle of a collision increases with collision energy. In addition, at higher collision energy electron detachment may occur and result in the formation of undetectable neutral species. To avoid poor collection efficiency due to these effects, the SRM experiments were conducted at low collision energy conditions. Figure 3.10 illustrates the improved selectivity of SRM over SIM. These data were obtained by the extraction and oxidation of a 2 ml plasma sample containing 40 n g Figure 3.9 96 800000 I —I f I ' I I I - (l ' E 0.6 mtorr argon 600000 " ‘ § .- ° 0.4 mtorr argon I,” ’ . 1.1 mtorr argon 7:: 400000 - - :1 p 200000 " ‘ I o \E‘. i : 0 . r . ‘9 0 15 20 collision offset energy (eV) The abundance of the (M-I-IF)- ion, m/z 310, from CAD of the molecular anion of 11,17-keto dexamethasone at different collision offset energies and collision gas pressures. 97 of internal standard and a dexamethasone concentration of 0.6 ng/ml. About 20 pg of oxidized dexamethasone extracted from the plasma sample and 0.8 ng of the oxidized isotopically labelled dexamethasone were injected onto the column for each analysis. The sum of the selected ion currents, due to the M- and (M-HF)- of 11,17-keto dexamethasone and the 11,17-keto analog of the 13C6-2H3-dexamethasone, is shown in Figure 3.10a. The sum of the selected reaction current obtained by SRM of the M- of 11,17-keto dexamethasone and of the 11,17-keto internal standard is shown in Figure 3.10b. In Figure 3.10b, all the interfering ion current is eliminated. The only signal present is that from the 11,17-keto dexamethasone and its 16B-CH3 epimer. The absence of background makes the determination of the peak height or the integration of the peak area straightforward and increases the precision of replicate analyses. Determination of the baseline position for peak height or area measurement from the SIM data in Figure 3.10a is more arbitrary which could lead to poorer precision than in SRM. The calibration curve obtained by SRM, shown in Figure 3.1 la is very similar to the calibration curve obtained by SIM that was shown in Figure 3.5a. The limits of detection of dexamethasone in plasma by the two techniques appear to be similar. The LOD of dexamethasone in plasma by SRM was calculated as 0.31 ng/ml from the y— intercept of the calibration curve and the standard deviation of the blank. It is likely that the LCD can be decreased by decreasing the amount of internal standard that is added. As found for SIM, the precision with which the blank can be measured, in addition to the small slope of the calibration curve, limit the sensitivity of the SRM assay. An additional experiment was done to demonstrate how the LOD for dexamethasone in plasma can be decreased by the addition of less internal standard. Preparation of these standards, in which ten times less internal standard was added, is described in part II of this chapter. The calibration curve obtained from these standards by SRM is shown in Figure 3.11b. Based on this calibration curve and the standard deviation obtained upon measuring the blank (0.005, 11 %RSD), an LOD of 0.05 ng/ml was calculated! This is three times Figure 3.10 98 8) m/z 310+330+319+339 SIM 100‘ ' 80 60- 40‘ 20- Q If:'I'I'TWfiwrfiiI'I'I'I‘I*T‘fi'I'I‘I'I‘I‘ t) 11 12 13 14 _15 16 b) 100- m/z 310+319 SRM r 80‘ M:->(M-HF)= 60-1 40, 20- w 0*1* 'U'IfI‘V'T‘I'V‘T'I‘U‘ ' ‘U'I'U'V'l' 'r 10 11 12 13 t 13 15 1‘6 Comparison of selected ion current profiles obtained by a) GC- ECNI/MS using SIM, with those obtained by b) GC-ECNI/MS/MS with SRM, for' the detemination of dexamethasone in a real plasma sample. 99 0.7 W I ' l ' F + A 0.6 - ‘33 ’ 1 m 0.5 - ' N I E 0.4 - - : I- a I: § 0.3 - '1 m N 0.2 - ‘ E 01 _ y=2.9le-2+3.43e-2x RA2=0.997 _ 0 0 1 I .L l 1 I 1 q 0 5 10 15 20 ng dexamethasone A 0.3 r O\ '_I m E :1 0.2 - ' \ § 1 m N E 0.1 - ' y = 5.25e-2 + 0.293;: RA2 = 0.985 0 0 _ , . r . . l . . I . . r r . 0.0 0.2 0.4 0.6 0.8 1.0 ng dexamethasone Figure 3.11 Calibration curves for determination of dexamethasone in plasma b GC-ECNI/MS/MS with SRM using a) 35 ng, and b) 3.5 ng of 1 C6-2H3-dexamethasone internal standard. 1 0 0 lower than the LOD calculated from the calibration curve obtained by SIM of these standards. The improved LOD is the result of a lower standard deviation and a steeper sloped calibration curve by SRM. Some of the same plasma samples from patients undergoing the dexamethasone suppresion test analyzed by SIM were analyzed by SRM. These results are shown in Table 3.3. The concentrations found using the direct probe sample introduction are also shown in Table 3.3. The concentrations determined for these five samples are similar by all three modes of analysis. Each sample (1.5 ml of plasma) was processed once. Also shown in Table 3.3 are the concentrations determined for the duplicate 1 ml plasma standards that were "spiked" with 2.09 ng of dexamethasone. The concentrations are reported in ng/ml of plasma +/- one standard deviation. The number of replicate instrumental analyses (n) is three unless indicated otherwise. The standard deviations shown represent the precision of the instrumental analysis. Table 3.3 Comparison of concentrations of dexamethasone obtained by GC- ECNI-SIM, GC-ECNT-SRM, and DIP-ECNI-SRM. Sample GC-SIM GC-SRM DIP-SRM P2 0.63 :1: 0.07 (n=2) 0.76 i 0.01 0.68 i- 0.03 (n=4) P5 0.62 :l: 0.09 0.69 :l: 0.02 0.35 i 0.01 P11 3.2 i 0.4 (n=2) 4.0 i 0.4 4.4 i 0.4 P18 3014 (n=2) 36:1: 3 35 i 1.0 P3 0.18 i 0.01 (n=4) 0.29 i 0.01 0.15 i 0.01 7A* 2.1iO.1 1.9i0.1 2.11:0.05 78* 1.8 i- 0.1 1.82 i 0.04 *duplicate spiked plasma sample (2.09 ng/ml), measured in triplicate The calibration curve obtained using the DIP with SRM was similar in appearance to the curves obtained by GC-SIM and GC-SRM. The concentrations determined by DIP-SRM are comparable to those found by GC-SIM and GC-SRM. This indicates that the added selectivity of MS/MS can allow the elimination of the chromatographic inlet. This will greatly reduce analysis time. For instance, to obtain the 1 O 1 results shown in Table 3.3, required five hours using the DIP. To obtain the GC-SRM results, 15 hours was required. Other researchers have noted that the chromatographic step is necessary in negative ion SRM when a general derivatization method is used to prepare an electrophilic derivative of the sample. This is because the electron capture responses of the matrix components are enhanced as well. Without sufficient chromatographic separation, the low level of analyte present in the ion source is unable to effectively compete with the matrix for the pool of thermal electrons. However, oxidation selectively converts the analyte to an electrophilic analog without greatly enhancing the response of the matrix. Therefore, introduction of an oxidized sample by direct probe is viable because the matrix is more transparent to electron capture and does not hinder the electron capture ionization of the analyte. C. Conclusion The use of selected reaction monitoring for the detection of oxidized dexamethasone in a plasma matrix yields similar results to those obtained through the use of selected ion monitoring. The additional selectivity of SRM allows the direct probe sample inlet to be used. However, when the GC inlet is used, SRM offers no significant advantage over SIM. This is due to the high level of selectivity already achieved through the oxidation and the electron capture negative ionization, as well as the chromatographic inlet. SAP?!" REFERENCES K. Minagawa, Y. Kasuya, S Baba, G. Knapp, and J. Skelly, J. Chromatogr., 343, 231-237, 1985. M. T. Lowy and H. Y. Meltzer, Biol. Psychiatry, 22, 373-385, 1987. M. Oehme, D. Stock], and H. Knoppel, Anal. Chem, 58, 554-558, 1986. G. R. Her and J. T. Watson, Anal. Biochem, 151, 292-298, 1985. American Chemical Society Committee on Environmental Improvement and Subcommittee on Environmental Analytical Chemistry, Anal. Chem, 52, 2242- 2249, 1980. M. J. Green and B. N. Lutsky, "Anti-Inflammatory Steroids" in CRC Handbook of Cardiovascular and Anti-Inflammatory Agents, M. Verderame, ed., CRC Press, Inc., Boca Raton, FL, pp. 1-26, 1986. F. W. McLafferty, ed., Tandem Mass Spectrometry, Wiley, New York, 1983. M. Dawson, C. M. McGee, P. M. Brooks, J. H. Vine, and T. R. Watson, Biomed. Environ. Mass Spectrom, I 7, 205-211, 1988. 102 CHAPTER 4. INVESTIGATION OF AN OXIDATION BASED ASSAY FOR 6B-HYDROXYCORTISOL IN URINE; COMPARISON WITH A CONVENTIONAL GC-MS METHOD 1. Introduction This chapter describes the investigation of an oxidation-based, GC-ECNI/MS assay for urinary 6B—hydroxycortisol. Results from this assay are compared to results obtained by a conventional GC-MS method based on formation of the methoxime- trimethylsilyl ether derivative of 6B-hydroxycortisol. The use of the direct inlet probe (DIP) is investigated for the ECNI/MS analysis to determine whether the overall selectivity provided by the oxidative treatment of the sample and the mode of ionization is sufficient to allow elimination of the chromatographic inlet. A. 6B-Hydroxycortisol and Enzyme Induction The compound 6B-hydroxycortisol (4-pregnene-6B,1lB,17a,20-tetrol-3,20-dione) is a minor metabolite of cortisol that is excreted in urine in free form (not conjugated with sulfate or glucuronate). Normal adults excrete between 200 and 400 ug of 6B- hydroxycortisol per day. This amount represents only 2% of the total weight of cortisol metabolites excreted in urine each day. Therefore, selective methodology is needed to detect the relatively small quantity of 6B-hydroxycortisol within the urine matrix. Unlike the other pathways for the metabolism of cortisol that involve reduction of the 4-ene-3-one moiety followed by conjugation with glucuronic acid, the formation of 6B-hydroxycortisol (6B-OHC) occurs through oxygenation by hepatic cytochrome P-450 microsomal enzymes. Upon induction of the P-450 enzymes, the excretion of 6B-OHC increases markedly. Enzyme induction'is defined as a state in which the rate of enzyme 103 l 04 protein synthesis is stimulated to a level above the normal rate of synthesis. Enzyme induction can be triggered by exposure to a variety of xenobiotic substances. Some examples include drugs, such as phenobarbital, and pesticides, such’as DDT (1). PBBs (2), polycyclic aromatic hydrocarbons (1), and styrene (3) also have been reported to induce microsomal enzymes. The measurement of urinary 6B-OHC has been proposed as anon-invasive means to monitor microsomal enzyme induction (1). The measurement of 6B-OHC has also been used to assess adrenocortical function in health and disease (4), and to assess the abnormality of cortisol metabolism in cancer patients (5), and as a diagnostic test for hypercortisolemic states (6). The major microsomal steroid hormone 6B-hydroxylase enzyme was recently identified as P-450N1: (7). These authors suggest that the level of urinary 613- hydroxylated steroids is likely to reflect the activity of only one or a limited number of closely related P-450 enzymes. The 6B-hydroxylated steroids are likely to be useful indicators of the metabolism of xenobiotic substances that are principally metabolized by P-450NF. These would include many compounds such as dihydropyridine calcium channel blockers, quinidine, erythromycin, benzphetamine, and aldrin. B. Methods for Determination of Urinary 6B-Hydroxycortisol There are many analytical methods available for the determination of 6B- hydroxycortisol in urine. The older methods, based on rigorous isolation of 6B-OHC with detection by a colorimetric technique (8), have been replaced by HPLC and RIA methods (9,10). A study comparing results obtained by an HPLC method and by RIA found a good correlation between the two methods (11). A line with a slope of 0.89 and a y-intercept of 22.03 was obtained by plotting the amount of 6B-OHC found by RIA on the x-axis against the amount found by HPLC. At higher levels of 6B-OHC, the RIA 1 05 values tended to be higher than the HPLC values. The higher levels of 6B-OHC were found in samples taken from subjects that had taken the drugs rifampicin and antipyrine in order to cause enzyme induction. The higher values found by RIA could be an indication of cross-reactivity with other substances that also increase upon enzyme induction. At the lower levels of 6B-OHC, the HPLC values were generally higher than the RIA values. This could be an indication of a co-eluting interference in the HPLC assay. At lower levels of 6B-OHC, interference from co-eluting material would become more pronounced. Other methods mentioned in the literature for the determination of 6B-OHC include GC-FID and GC-MS with the formation of the MO-TMS derivative (12). A GC- ECD assay based on oxidation of 6B-OHC (13) is also mentioned. These methods have not been used extensively. The GC-ECD assay has not found extensive use. Criticism of the method included the possibility that other components of the urine matrix could be oxidized to the same product as 6B-hydroxycortisol. Also, the poor chromatographic resolution ' available with the packed column might not separate the oxidized 6B-hydroxycortisol from other similar oxidized components that have an electron capture response. C. Proposed ECNI-MS Method for Determination of 6B-Hydroxycortisol The basis for our proposed assay for 6B-hydroxycortisol is the chemical oxidation of 6B-hydroxycortisol to an electrophilic product. Figure 4.1 shows the oxidative conversion of 6B—hydroxycortisol to 4-androstene-3,6,l1,17-one. The 4-ene-3,6-dione moiety incorporated into the molecule is primarily responsible for the high electron capture response. Cleavage of the C17 sidechain makes the 6B-hydroxycortisol analog thermally stable during GC analysis. 1 o 6 CHz-OH c=o HO "OH oxidation OH I IIIIIIIUIFTFIIITIIIIIT—T‘ITIIUIIITTI 2802852932953D3530315320325330335340345350 Figure 4.1 a) Oxidative conversion of 6B—hydroxycortisol to electrophilic product. b) ECNI mass spectrum of oxidized 6B-hydroxycortisol. l 0 7 The ECNI mass spectrum of oxidized 6B-OHC is also shown in Figure 4.1. The only peak in the spectrum, at m/z 314, represents the molecular anion. The additional selectivity of the mass information provided in ECNI-MS analysis eliminates the GC-ECD problem of interference from co-eluting compounds having electron capture response. However, as in the GC-ECD assay, the criticism of an ECNI- MS method lies in the question of whether the amount of 4-androstene-3,6,11,17-one formed upon oxidation of an urine extract truly represents the amount of 6B- hydroxycortisol in that extract. A literature search was done to help answer this question. According to the literature that was surveyed, 6a-hydroxycortisol is the only compound, of those normally excreted in human urine in amounts comparable to 6B-hydroxycortisol, that may be oxidized to 4-androstene-3,6,l1,17-one. The compound 6a-hydroxycortisol normally comprises about 5-10% of the total 6-hydroxycortisol that is excreted (8). The measurement of the total 6-hydroxycortisols has also been used as an index of microsomal enzyme induction (14). 11. Experimental and Results This section contains the details of the method development and the results obtained for the determination of 6B-hydroxycortisol in urine. Three studies are discussed here. These include a preliminary study, a study using urine from control and phenobarbital treated guinea pigs, and a study using human urine from a volunteer before and after receiving niacin treatment to lower cholesterol. 108 A. Preliminary Study This study was done to assess the feasibility of oxidation and ECNI-MS for the detection of 6B-hydroxycortisol in human urine. The ECNI mass spectra of other oxidized components of urine could produce a response at m/z 314. This preliminary study was also done to determine the extent of interference at m/z 314 from the oxidized urine matrix. The sample preparation, the ECNI parameters, and the results are discussed. 1. Extraction of Urine A large volume (120 ml) of human urine was used for this study to ensure detection of 6B-hydroxycortisol as well as other components of the urine matrix. The steroids were extracted from the urine by reverse-phase, solid-phase extraction. After passing the sample through the C18 solid-phase extraction (SPE) column, the column was washed with 5 ml of distilled water. The steroids were eluted with 3 ml of methanol. Methanol elutes both conjugated and free steroids. The methanol was evaporated and the residue reconstituted in 1 ml of water. The water was extracted three times with 1-2 ml of ethyl acetate to isolate the free steroids from the conjugated steroids. The combined ethyl acetate extracts were divided into three portions. One of the portions was used for the following oxidation. 2. Oxidation The ethyl acetate extract was reconstituted in 1 ml of CH2C12 for the oxidation. An excess of the oxidation reagent, pyridinium chlorochromate (PCC) finely ground up with celite, was added to the solution. Sodium acetate was added to buffer acid liberated 1 O 9 during the oxidation. The reaction was stirred at room temperature for 4.5 hours. The excess oxidation reagent was removed by passing the reaction mixture through a silica (500 mg) SPE column. Four milliliters of ethyl acetate were used to elute the oxidized 6B-hydroxycortisol. The ethyl acetate was evaporated and the residue reconstituted in 100 111 ethyl acetate. 3. GC-ECNI/MS and GC-EI/MS Analysis of Oxidized Urine Extract The oxidized urine extract was analyzed by GC-ECNI/MS and by GC-EI/MS. The instrument used for these analyses was an I-IP5985 quadrupole mass spectrometer that had been modified in this laboratory to allow negative ion detection. Preliminary work with standards of oxidized 6B-OHC indicated that sensitivity increased with increasing source temperature. The total ion current (TIC) from oxidized 6B- hydroxycortisol was measured at source temperatures of 100°C, 150°C, and 200°C. The total ion current at a source temperature of 200°C was forty times greater than the TIC at a temperature of 100°C. A source temperature of 200°C was used for this study. Methane was used as the modifying gas. The pressure was adjusted using an automatic pressure controller (APC) valve that was feedback-regulated by the ionization gauge. A reading of 2.7 x 10 - 4 on the ionization gauge situated below the ion source housing was maintained by the APC valve. This ion gauge reading corresponded to an ion source pressure of about 0.3 torr. The GC-MS was equipped with an on-column injector and a "megabore" DB-5 capillary column (25 m x 0.53 mm). The column effluent was introduced to the mass spectrometer via an open-split interface. About one-half to one-fifth of the column effluent was transferred to the MS by this interface; the exact amount depended on the flow rate of the carrier gas through the column. At higher column flow rates, a greater proportion of the effluent is split out because the flow through the transfer line into the l 1 0 mass spectrometer is limited by the narrow inner diameter of the tubing (0.11 mm). The column was operated at a head pressure of 5 psi. This corresponded to a flow rate of about 5 ml/min. For the GC—EI/MS analysis, conventional mass spectrometer conditions were used (ion source temperature = 200°C, 70 eV). The results of the analysis of the oxidized urine extract by GC-EI/MS are shown is Figure 4.2a. The use of the non-selective electron impact ionization gives an indication of the complexity of the sample because all of the sample components are ionized. The peak corresponding to oxidized 6B- hydroxycortisol (indicated with an arrow) indicates that this is only a minor component of the sample. The results of the analysis of the oxidized urine extract by GC—ECNI/MS are shown in Figure 4.2b. Under ECNI conditions, oxidized 6B-hydroxycortisol is respresented by one of the major peaks in the reconstructed total ion current (TIC) chromatogram. Very few of the other components in the matrix give an appreciable response. Comparison of the reconstructed total ion current chromatograms in Figure 4.2 demonstrates the high degree of selectivity for 6B-hydroxycortisol that is imparted by the oxidative treatment of the sample and subsequent analysis by ECNI. The increase in absolute sensitivity of detection is also evident by comparison of the 100% TIC for the molecular ion of oxidized 6B-hydroxycortisol by El and by ECNI, shown in Figure 4.23 and 4.2b, respectively. . 4. Investigation of Sample Introduction by Direct Inlet Probe for ECNI/MS The reconstructed mass chromatogram at m/z 314 from the GC-ECNI/MS analysis of the oxidized urine extract is also shown in Figure 4.2b. This mass chromatogram indicates that the only major response at m/z 314 is from the molecular Figure 4.2 111 ) 100°/.=4328 “0 "'I'I‘I“'1"‘I"'ITTTI"'1"'17"I"‘I'I'IH'IH‘ITTT 4 5 6 7 6 9 10 11 12 13 14 15 16 17 100%=11 m/z314 ITIV‘"I'1_"T"‘I'JTT“‘I“‘Ifi'1**‘7“‘l"'I"‘I"'1" 4 5 6 7 6 9 10 11 12 13 14 15 16 17 T1me(minutes) ) J100%=10384 TIC A l LMAA A; A T‘TTT'TI‘ITTT' 17"I'Ifin'l‘7‘l7ffijfi'.‘T'In'lh7TT‘ 4 5 6 7 6 O 10 11 12 13 14 15 16 17 1 100%-3370 111/2314 11 -. 1 1- 1 _1' 1 I I‘ I ‘1 'IT' I "I "1"‘Ifi71fl'17‘l‘ ‘ITT 4 5 6 7 8 9 10 11 12 13 14 15 16 17 Time (minutes) Reconstructed total ron current chromato chromatogram at m/z gram and mass 314 from a) GC-El/MS and b) GC-ECNI/MS analyses of an oxidized urine extract. l 12 anion of oxidized 6l3-hydroxycortisol. Because no other compounds in the sample matrix produce an appreciable response at m/z 314, it may be possible to use the direct inlet probe (DIP) and selected ion monitoring to determine oxidized 6B-hydroxycortisol in urine. This would simplify the analysis and reduce the time required. It was found that the best signal-to-background (s/b) ratio could be achieved by rapidly heating the DIP. Rapid heating of the DIP also reduces the possibility of sample decomposition. When the probe is heated rapidly, the rate of desorption is greater than the rate of decomposition (15). The maximum heating rate allowed by the data system was 30°C/min. To achieve a faster heating rate, the heating cable was disconnected from the probe and the initial probe temperature was set to 280°C through the data system. Then, after insertion of the probe into the ion source, the cable was reconnected. A heating rate of about 200°C/min can be achieved by this procedure. The desorption profile shown in Figure 4.3 was obtained for the analysis of the oxidized urine extract by ECNI with SIM, using rapid heating. The use of the slower heating rates available through the data system resulted in much broader and slower desorption profiles. A sharp, non-tailing desorption profile is much easier to integrate, leading to better precision in quantitative analyses. 5. Conclusion This study indicated that GC-ECNI/MS provides good selectivity and sensitivty for the detection of the oxidation product of 6B-hydroxycortisol within the urine matrix. The feasibility of using this methodology to determine urinary 6B-hydroxycortisol was demonstrated. 113 SIM at m/z 314 Area = 76400 Time (minutes) Figure 4.3 Selected ion current profile obtained by DIP-ECNI/MS analysis of an oxidized urine extract. The molecular anion of oxidized 6B- hydroxycortisol corresponds to m/z 314. 114 B. Determination of 6|3-Hydroxycortisol in Guinea Pig Urine. A study was done in which the urines of a control and an "induced" guinea pig were analyzed for 6B-hydroxycortisol. The "induced" guinea pig was given phenobarbital, which is known to cause P-450 enzyme induction. There was indication in literature that guinea pigs excrete 6B-hydroxycortisol and that the amount excreted increases upon P-450 enzyme induction with phenobarbital (16). This study was undertaken to determine whether the oxidation methodology for the determination of 613- hydroxycortisol would indicate that enzyme induction had occurred in the phenobarbital treated animals. The guinea pig urines were also analyzed by the conventional GC-MS method for steroid analysis, in which the methoxime-trimethtyl silyl ether derivatives are made. The results of the conventional method were compared to the results obtained by oxidation and ECNI-MS. 1. Sample Preparation The protocol followed for the sample preparation is illustrated by the scheme shown in Figure 4.4. Three-quarters of the extract was used for the MO-TMS procedure because it was anticipated that the EI/MS detection might not be as sensitive as the ECNI detection. The control and the induced samples were processed in triplicate. The internal standard, 4-pregnene-6B-ol-3,l1,20-one, was chosen because: 1) this compound is oxidized to an electrophilic product that was not found in oxidized guinea pig urine by a preliminary analysis; 2) the compound undergoes the reactions to form an MO-TMS derivative; 3) the compound is structurally similar to 6B-hydroxycortisol and ought to behave similarly throughout the isolation procedures, although it is somewhat less polar than 6B-hydroxycortiso1; and 4) it was not found in guinea pig urine by a preliminary analysis using the MO-TMS method. 115 SAMPLE PREPARATION 2% of total daily urine + 680' ng internal standard 1 C18 Sep-PakTN extraction solvent extraction I 34 75% aliquot ' 25% aliquot Methoxyamine, 60°C, 1hour PCC oxidation, 5 hours Trimethylsilylimidazole, 100°C, 4 hours silica Bond-Elutm 1 Sephadex LH-20 column Sep-Pakm is a trade name of Waters Associates Bond-Elutw is a trade name of Analytichem International Figure 4.4 Sample preparation scheme for the determination of 6B- hydroxycortisol in urine by MO-TMS derivatization and by oxidation. 1 1 6 Standards containing internal standard and 6B-hydroxycortisol were prepared in duplicate for the construction of a calibration curve. The procedures applied for the analysis of the guinea pig urine were also used to process the standards. 2. Instrumental Analysis These samples and standards were analyzed with three different instruments. Analysis of the MO-TMS samples by GC-EI/MS with SIM was initially performed on the HP5985 mass spectrometer. Selected ion monitoring was used to monitor the ion current from the molecular ion and the (M-OCH3)+ ion of 6B—hydroxycortisol MO-TMS, and the molecular ion and the (M-CH3)+ ion of the internal standard. These ions corresponded to m/z values of 724, 693, 474, and 459, respectively. The resolution of the GC capillary column (12 m x 0.2 mm methyl silicone, SP2100) was not sufficient to resolve the 6B-hydroxycortisol MO-TMS derivative completely from sample constituents that gave response at the same m/z values as the molecular ion and (M-OCH3)+ ion of 6B-hydroxycortisol MO-TMS. The results from the analysis of a derivatized guinea pig— urine obtained with the I-IP5985 are shown in Figure 4.5a. The retention time of the 6B- hydroxycortisol MO-TMS derivative is indicated with an arrow. The guinea pig urine samples were re-analyzed later using a JEOL AX505 double focussing mass spectrometer. A HP5990 gas chromatograph was directly interfaced with this mass spectrometer. Better chromatograpic resolution was obtained using this system and a new 30 m x 0.32 mm DB-l capillary column. Splitless injection was used to introduce the samples. The temperature program was started at 50°C to condense the solvent and the sample components. The oven was ramped up to 260°C at 35°C/min. and then ramped up to 290°C at 4°C/min. The ion currents at m/z 724, 693, 474, and 459 were monitored using magnetic field switching selected ion monitoring. The collector slit (the slit just prior to the electron multiplier) was opened up to 200 um to maximize E | ECNI Figure 4.5 117 1 _ l m/2693 l I l T T l ‘ I l l l l : m/z459,474 - internal - standard ‘T"'1"7T"‘l"‘f“'l"‘l'TT 12 13 14 15 16 17 18 19 m/z 314 rn/z 342 internal standard Time (minutes) Selected ion current profiles obtained from a) GC-EI/MS analysis of MO-TMS derivatized guinea pig urine sample and b) GC- ECNI/MS analysis of oxidized guinea pig urine sample. 118 the detection of these ions. An accelerating voltage of 3 kV and post accelerating voltage of 10 kV were used. Due to sensitivity problems experienced with negative ion detection by the HP5985, the analyses of the oxidized samples by GC-ECNI/MS were performed on the JEOL HXl 10 mass spectrometer. At the time of these analyses, the HXllO was interfaced with the HP5990 GC via a single stage jet separator. The samples entered the separator via a "megabore" DB-l capillary column (15 m x 0.53 mm) with a carrier gas flow rate of 10-12 ml/min. Make-up gas was fed into the separator for a total flow rate of about 30 ml/min. An ion source temperature of 200°C was used for these analyses. Methane was used as the modifying gas. The HXllO was not capable of performing magnetic field switching, so electric field switching was used for the selected ion monitoring. Magnetic field switching SIM is preferred over electric field (EF) switching SIM because the latter discriminates against the ions of higher masses that are monitored within an analysis. This is because the accelerating voltage is decreased to pass the higher mass ions during EF switching SIM. The data shown in Figure 4.5b were obtained during EF switching SIM analysis of an oxidized guinea pig urine. DIP-ECNI/MS analyses of the oxidized guinea pig urines were performed on the HP5985. (The negative ion sensitivty had been restored to an acceptable level by various maintenance procedures.) The molecular anions of oxidized 6B-hydroxycortisol and the oxidized internal standard, at m/z 314 and m/z 342, respectively, were monitored. The ion source parameters were the same as those used for the preliminary study of oxidized human urine. The direct probe was heated rapidly to 250°C. A good desorption profile was obtained by rapidly heating the probe to 250°C. 119 3. Results and Discussion The results of the GC-EI/MS, GC-ECNI/MS, and the DIP-ECNI/MS analyses of the control and guinea pig urines are shown in Table 4.1. The values in Table 4.1 represent the total daily excretion of 6B-hydroxycortisol (pg/day). The number of analyses (n) for each sample was one unless it is indicated as otherwise. Table 4.1 Comparison of GC-EI/MS of the MO-TMS derivatived urine with GC—ECNI/MS, and DIP-ECNI/MS, of the oxidized urine, for the determination of 6B-hydroxycortisol excretion by control and induced guinea pigs. Sample GC-EI GC-ECNI DIP-ECNI control 1a 74 62 1 19 control 1b 61 48 95 i- 1 (n=2) control 1c 55 50 101 i 4 (n=2) mean 63 i 10 53 i 7 102 i 10 induced la 20 77 1 l 1 induced 1b 16 70 92 induced 1c 17 64 127 mean 18:2 701-6 110:18 The paired t-test was used to determine whether the three methods produced significantly different results for the determination of 6B-OHC in the control and the induced samples. The analysis of the control sample by GC-EI/MS and by GC-ECNI/MS did not yield significantly different results. The results found by DIP-ECNI/MS were significantly different from the results found by both GC-EI/MS and GC-ECNI/MS. This indicates that the oxidation method can provide results that are comparable to the conventional GC-MS method for the analysis of normal guinea pig urine, provided that the GC inlet is used. Unfortunately, the analyses using the DIP gave erroneously high 120 results. This would indicate that other compounds within the oxidized urine matrix are capable of producing a response at m/z 314 under ECNI conditions. For the induced sample, the results from the GC-EI/MS analysis were significantly different from the results of the ECNI/MS analyses. The GC- and DIP— ECNI/MS results were not significantly different by the paired t-test. This indicates that, in the induced state, the guinea pig may excrete compounds other than 613- hydroxycortisol that are capable of being oxidized to 4-androstene-3,6,l1,17-one. The regular t-test was applied to test for a significant difference between the control and induced samples. The difference between the control and induced samples by the GC-ECNI/MS method was significant at the 95% confidence level. The control and induced samples were also significantly different by the GC-EI/MS method even at the 99% confidence level. By DIP-ECNI/MS analysis, the difference between the control and induced samples was not significant. It is interesting that the results found 1;; the induced sample by the GC-EI/MS method are significantly lower than those for the control sample. This contradicts the report that the amount of 6B-hydroxycortisol excreted by guinea pigs increases upon P- 450 enzyme induction with phenobarbital. Possible reasons for this disagreement could be due to the methodology used by the researchers who reported these results. The method used in their determinations was based on extraction and paper chromatography and the Porter—Silber reaction for colorimetric determination. A more recent study (17) has indicated that "6B-hydroxycortisol is not a useful index of hepatic microsomal enzyme induction in the guinea pig because 6B-OHC itself undergoes extensive metabolism in this species." The method for 6B-OHC determination used in this study was RIA. The metabolites of 6B-OHC were not identified, but it was suggested that the metabolites might contain a C21 carboxyl group, as they were acid extractable. It is very likely that 21—carboxy 6l3-OHC could be converted to 4-androstene—3,6,1 1,17-one by the PCC oxidation. This may explain the why the results found by the oxidation method for 1 2 l the induced urine were higher than the results obtained by the MO-TMS method. The relative standard deviations were 13-15% for the GC-EI/MS analyses, 8-13% for the GC-ECNI/MS analyses, and 15% for the DIP-ECNI/MS analyses. With an internal standard it should be possible to achieve better precision. The use of an internal standard that more closely resembles 6B-hydroxycortisol may help to improve the precision of the determination of 6B-OHC by these methods. Ideally, an isotopically- labelled analog should be used for the internal standard. C. Determination of 6B-Hydroxycortisol in Human Urine by GC-EI/MS, GC-ECNI/MS, and DIP-ECNI/MS The results of the study described above indicated that the guinea pig was not a good model for studying 6B-hydroxylation of cortisol. The goal of this proposed methodology is to provide a reliable and relatively simple means to determine 6B-OHC in human urine. Another study was undertaken to determine the reliability of the oxidation method for the determination of 6B-hydroxycortisol in human urine. The excretion rate of 6B-OHC in normal human subjects is 200-400 ug/day. The samples used in For this study twenty four-hour urine samples were collected before and after the volunteer subject underwent niacin treatment to lower blood cholesterol levels. These urine samples were analyzed for 6B-hydroxycortisol by the oxidation method with ECNI/MS detection and the conventional MO-TMS method with GC-EI/MS analysis. The results obtained by the different methods are compared. 1. Sample Preparations The preparation of these samples was the same as that for the guinea pig urines, with a few changes. The volume of the human urine used for these analyses was 2 ml. This represented about 0.1 to 0.2% of the total amount of the 24-hour urine samples. The 122 compound 6B-hydroxyprednisolone was used as the internal standard. This compound is non-endogenous and is identical to 6B-OHC except for the presence of a double bond in the C1 position. Attempts to synthesize a deuterated analog of 6B-OHC for use as the internal standard were not successful. (A description of these experiments and the results are included below.) The ethyl acetate extraction following the C18 solid-phase extraction of the urine was eliminated. The oxidation was conducted at 60°C instead of room temperature. Experiments described in Chapter 2 indicated that the rate of oxidation can be increased by heating the reaction. The reaction time used was 4.5 hours. A comparison of the yield of 4-androstene—3,6,ll,l7-one from urinary 6B- hydroxycortisol indicated that the yield was about the same after 2, 3, and 5 hours of oxidation. Removal of the excess PCC for this study was accomplished with a 200 mg silica SPE column. Previously, 500 mg columns had been used. After loading the sample onto the column, the column was washed with 3 ml of CH2C12, and the oxidized 6B-OHC was eluted with 2 ml of ethyl acetate. Nearly all of the oxidized 6B-OHC is eluted in this fraction by 2 ml of ethyl acetate. With larger volumes of ethyl acetate, the PCC starts to elute from the column. 2. Deuteration Experiments The goal of these experiments was to incorporate three deuterons by exchange of the alpha hydrogens of the C3 ketone. In the first experiment, excess solid sodium methoxide was added to a solution of 6B-OHC in CH3OD. The reaction mixture was stirred at room temperature in a capped vial for seven days. The mixture was then diluted with CH2C12. The CH2C12 layer was isolated and washed with water to remove the sodium methoxide. The CH2C12 was dried with anhydrous sodium sulfate to remove traces of water. Analysis of the reaction mixture by GC-EI/MS revealed the presence of a compound that could have been hexadeutero l7-keto 6B-OHC. 123 In the second experiment, 6B-OHC was dissolved in CH3OD and a piece of sodium metal was added directly to the solution. The reduction of the Na(s) proceeded spontaneously resulting in the in situ production of sodium methoxide and deuterium gas. The methoxide ion is a strong base and is capable of pulling off acidic hydrogens, including alpha-keto hydrogens. After 1 hour, an aliquot of the reaction mixture was removed and diluted with CH2C12. The CH2C12 layer was isolated, washed with water, and dried with anhydrous sodium sulfate. After 7 hours another aliquot of the reaction was extracted, washed, and dried. No intact 6B—OHC was recovered in either of these aliquots. Apparently, the compound is degraded in the presence of such a strong base. Another method that was tried involved the use of basic alumina (18). Basic alumina (8.2 g) having Brockmann I activity was incubated with D20 (0.25 ml) for 2 1/2 hours. About one quarter of the alumina was poured into a pasteur pipette plugged at the bottom with glass wool. The column was wetted with acetonitrile and 0.5 ml of a solution of 6B-OHC in CH3OD was loaded onto the column. The exchangeable hydrogens were to have been replaced with deuterium by the time the compound eluted from the column. Acetonitrile was used as the eluent. Several fractions of 3.5 ml were collected from the column. The fractions were analyzed by HPLC to see if they contained any 6B-OHC. The first fraction contained a degradation product of 6B-OHC, 17-keto-6B-OHC. No 6B-hydroxycortisol was recovered from the column. The identity of the degradation product was determined by GC-EI/MS of its MO-TMS derivative. No deuterium was incorporated into this molecule. If a less polar eluent had been used, the 6B-OHC would have remained on the column for a longer time. This might have allowed more time for the exchange to take place. The labile C17 sidechain must be protected to prevent degradation during the exchange reaction. This could be accomplished by the formation of the bis- methylenedioxy derivative (19). 124 3. Instrumental Analysis Two different mass spectrometers were used for the analysis of these urine samples. The JEOL AX505 was used for the GC-EI/MS analysis of the MO-TMS samples. The mass spectrometer conditions were the same as those described for the analysis of the MO-TMS of the guinea pig urine samples. The AX505 was also used for the DIP-ECNI/MS analysis of the oxidized urines. With this instrument, very high ion source temperature is required to efficiently ionize oxidized 6l3-OHC. Methane was used as the modifying gas at a pressure reading of 1.2 x 10-5 on the penning gauge. Selected ion monitoring by magnetic field switching was used to detect the molecular anions of oxidized 6B-OHC and oxidized 6B-hydroxyprednisolone, at m/z 314 and m/z 312, respectively. A good desorption profile was obtained on this instrument by heating the DIP from 30°C to 100°C at 32°C/min. The Finnigan TSQ70 triplequadrupole mass spectrometer was used for the GC- ECNI/MS analysis of the oxidized urine samples. The TSQ70 was operated as a conventional scanning mass spectrometer for these analysis. The negative ion sensitivity of this instrument is superior to that available with any of the other instruments. Parameters for the ECNI analysis were roughly optimized. Rigorous optimization was not necessary because the sensitivity was found to be adequate under the commonly used conditions for ECNI on this instrument. 125 4. Results and Discussion The results for the determination of 680HC in these urine samples by the MO- TMS and oxidation methods are compared in Table 4.2. The number of replicate analyses, n, was three, unless indicated otherwise. The values are reported i the standard deviation of the triplicate analyses. Table 4.2 Comparison of GC-EI/MS of MO-TMS derivatives with GC- ECNI/MS and DIP-ECNI/MS of oxidized urine samples for the determination of the total daily excretion of 6B-hydroxycortisol by a 49-year old male. Sam le GC-EI GC-ECNI DIP-ECNI control 1a 340 i 10 290 i- 10 920 i 190 (n=4) control 1b 380 i 30 345 i 20 niacin 1a 345 :l: 7 _ 280 i 20 niacin 1b 300 i 8 niacin 2a 377 (n=l) niacin 2b 371 (n=1) The paired t-test was applied to test for a significant difference between the oxidation GC-ECNI/MS method and the MO-TMS GC-EI/MS method for the determination of 6B-hydroxycortisol. The difference between the two methods was not significant even at the 98% confidence level. As found in the analysis of the guinea pig urines, the values determined using the DIP were much higher than the values found using the GC inlet. More rigorous separation of the 6B-hydroxycortisol from the other components in the urine samples appears to be necessary in order to use the DIP for sample introduction. Treatment with niacin does not appear to alter the amount of 6B-hydroxycortisol excreted per day. The amount of 6B-hydroxycortisol excreted per day as determined by the oxidation method and the MO-TMS method fell within the expected normal range of ZOO-400 ttg/day. 126 III. Conclusions The results described in this chapter have demonstrated that oxidation can be used to selectively enhance the ECNI-MS detection of 68-hydoxycortisol in human urine. The results obtained by this oxidation method do not differ significantly from the results obtained by the GC-MS reference method based on the formation of the MO-T‘MS derivative. The use of the DIP for the ECNI-MS analysis tended to give errroneosly high results. The procedure used to isolate the 6B-hydroxycortisol from the urine samples was not very selective for 6B-hydroxycortisol. Implementation of more selective isolation of 68-hydroxycortisol might remove matrix components that interfere with the analysis by DIP-ECNI. The use of oxidation and ECNI-MS for the determination of urinary 6B- hydroxycortisol offers some advantages over the MO-TMS method. The good selectivity of the methodology allows the use of simpler and faster GC separation. A "megabore" capillary GC column can be used to introduce the oxidized urine samples to a mass spectrometer. Also, only one product is produced by the one-step oxidation of 6B- hydroxycortisol. The MO-TMS derivatization requires more rigorous and lengthy GC separation. In addition to interference from sample components, syn and anti isomers of 6B—hydroxycortisol, produced in the formation of the methoximes, complicate the chromatographic separation. The MO-TMS derivatization is a two-step reaction. The use of a two-step reaction lengthens the sample preparation time and can lead to a mixture of products. The product of the oxidation reaction is quite stable and requires no special storage conditions. The MO—TMS derivative must be kept in a moisture free environment to avoid hydrolysis of the TMS ethers. 10. ll 12. l3. 14. 15. 16. 17. 18. REFERENCES A. H. Conney, Pharmacol. Rev., 19, 317-366, 1967. J. J. Vrabanac, Jr., Efi’ects of Exposure to Polybrominated Biphenyls on Urinary Steroid Metabolic Profiles in Man and Rat as Determined by Capillary GC and GC/MS/DS, Ph. D. Dissertation, Department of Pharmacology and Toxicology, Michigan State University, 1983. P. Dolara, M. Lodovici, M. Salvadori, G. Santoni, and G. Cademi, Ann. Occup. Hyg., 27, 183-188, 1983. S. B. Pal, Metabolism, 27, 1003-1011, 1978. E. M. Werk, Jr., J. MacGee, and L. J. Sholiton, Metabolism, 13,1425-1438, 1964. E. Voccia, P. Saenger, R. E. Peterson, W. Rauh, K. Gottesdiener, L. S. Levine, and M. I. New, J. C lin. Endocrinol. Metab., 48, 467-471, 1979. D. J. Waxman, C. Attisano, F. P. Guengerich, and G. P. Lapenson, Arch. Biochem. Biophys., 263, 424-436, 1988. A. G. Frantz, F. H. Katz, and J. W. Jailer, J. Clin. Endocrinol. Metab., 21, 1290- 1303, 1961. T. Ono, K. Tanida, H. Shibata, H. Shirnakawa, Chem. Pharm. Bull, 34, 2522- 2527, 1986. B. K. Park, J. Steroid Biochem, 9, 963-966, 1978. E. Gerber-Taras, B. K. Park, and E. E. Ohnhaus, J. Clin. Chem. Clin. Biochem, 19, 525-527, 1981. M. G. Homing, S. S. Lau, A. Hung, W. G. Stillwell, and R. M. Hill, J. Steroid Biochem, 5, 362 (abstract), 1974. J. Chamberlain, Clin. Chim. Acta, 34, 269-271, 1971. K. Thrasher, E. E. Werk, Jr., Y. Choi, L. J. Sholiton, W. Meyer, and C. Olinger, Steroids, 14, 455-468, 1969. R. J. Cotter, Anal. Chem, 52, 1589A-1606A, 1980. R. Kuntzman, M. Jacobson, W. Levin, and A. H. Conney, Biochem. Pharmacol., 17, 565-571, 1968. B. K. Park, M. R. Challiner, and S. Newby, J. Steroid Biochem, 18, 453-457, 1983. A. F. Hofmann, P. A. Szczepanik, and P. D. Klein, J. Lipid Res., 9, 707-713, 1968. 127 1 2 8 19. J. A. Edwards, M. C. Calzada, and A. Bowers, J. Med. Chem, 7, 528-530, 1964. CHAPTER 5. SUMMARY This research project involved development of new methodology for the mass spectrometric determination of corticosteroids in biological samples. Because of the selectivity and sensitivity associated with electron capture negative ionization (ECNI) mass spectrometry, this technique was investigated for enhancing- the detectability of corticosteroids. Corticosteroid determinations by ECNI have been very limited because the derivatization of this class of steroids with electrophilic moieties has been unsatisfactory due to the multiple hydroxy and ketone groups present in these types of molecules. A few types of steroids can be converted to highly electrophilic analogs by oxidation. Examples include steroids having the 4-ene-3-one-6-oxy moiety, the 1,4- ' diene-3-one-l l-oxy moiety, or the 5-ene-3-ol moiety. Previous work in this laboratory demonstrated the feasibility of using oxidation to enhance the electron capturing properties of corticosteroids drugs by incorporating C-11 and C-17 ketone moieties into these molecules. The major goals of this research project were: 1) the optimization of oxidation methodology for quantitative conversion of corticosteroids to their electrophilic oxidized analogs; 2) to validate this oxidation methodology for a) the determination of dexamethasone in plasma from patients undergoing the dexamethasone suppression test, and b) the determination of 6B-hydroxycortisol in urine; and 3) to investigate the feasibility of a faster, less selective method, such as direct probe, for introducing the sample into the mass spectrometer. Several Cr(VI) oxidants and various conditions for the oxidation of dexamethasone were studied to establish optimal production of 11,17-keto dexamethasone. A yield of 70% 11,17-keto dexamethasone could be achieved after 5.5 hours by PCC oxidation in CH2C12, with sodium acetate, at 60°C. Complete recovery of 129 1 3 0 the product from the reaction mixture could. be achieved using silica solid-phase extraction with ethyl acetate or 5% acetone-CH2C12 as the eluent. Some preliminary work indicated that the oxidation product of 6B-hydroxydexamethasone, the major metabolite of dexamethasone, also gave a high electron capture response. The oxidation methodology with ECNI-MS analysis for the determination of dexamethasone in plasma was validated by comparison with other methods. Concentrations of dexamethasone in plasma samples from patients undergoing the dexamethasone suppression test (DST) were determined using oxidation and GC- ECNI/MS with SIM. Aliquots of the same plasma samples were analyzed by GC-EI/MS and RIA methods in other laboratories. The results obtained by the oxidation method with ECNI-SIM detection did not differ greatly from the results obtained by the reference GC-EI-SIM method. The EI-MS method was based on the formation of the TMS-enol-TMS ether derivative of dexamethasone. All of the dexamethasone extracted from a 5 ml plasma sample was derivatized and concentrated to a volume of 5 111. All of the 5 111 was required for a single analysis by EI-MS. An aliquot rcpresenting only 2% of a 1.5 ml plasma sample was adequate for the ECNI-MS analyses. The relative standard deviations of the EI-MS method and the oxidation method with ECNI-MS were 16% and 10%, respectively. The concentrations of dexamethasone in the DST samples were also determined by RIA in nine different laboratories. The results obtained by the oxidation method generally fell within one standard deviation of the mean of the nine RIA determinations. The relative standard deviations associated with the nine RIA determinations were between 2% and 50% The use of GC-ECNI/MS/MS was investigated for the detection of oxidized dexamethasone in the plasma matrix. Selected reaction monitoring (SRM) was used to detect the (M-HF)- daughter ion of the molecular anion of 11,17-keto dexamethasone. This technique yielded similar results to those obtained through the use of selected ion Eh, 131 monitoring. If the GC inlet is used, SRM offers no significant advantage over SIM. This is due to the high level of selectivity already achieved through the oxidation and the electron capture negative ionization, as well as through the use of the chromatographic inlet. However, with the additional selectivity of SRM, the direct probe sample inlet can be used in place of the GC inlet. The plasma dexamethasone concentrations determined using the direct inlet probe were comparable to those obtained using the GC inlet with SIM and with SRM. Use of the direct probe can decrease analysis time to 3-5 minutes per sample. Typical GC analysis time is 15-20 minutes per sample. The only disadvantage associated with the oxidation method is the possibility that other components in the sample matrix could be oxidized to the same product as dexamethasone. The characteristic loss of HF from the molecular anion of oxidized dexamethasone differentiates oxidized dexamethasone from endogenous sample components because endogenous compounds do not contain fluorine. Interference from the metabolites of dexamethasone is not expected to be significant because the major metabolite of dexamethasone is 6B-hydroxydexamethasone. The oxidation product of 6B-hydroxy dexamethasone, 6,11,17-keto dexamethasone, does not have the same mass spectrum as 11,17-keto dexamethasone; therefore, no interference is expected. The oxidation product for the 20-dihydro metabolite of dexamethasone would be the same as that for dexamethasone. However, since this is a minor metabolite, its interference is not expected to severly limit this technique. The results that were obtained by the oxidation method for the determination of dexamethasone in the plasma samples were similar to the GC-EI/MS results. Therefore, components from the plasma matrix and dexamethasone metabolites do not appear to interfere with the determination of dexamethasone by the oxidation method with GC-ECNI/MS detection. The feasibility of using oxidation methodology to determine urinary 6B- hydroxycortisol was also demonstrated. The studies indicated that GC-ECNI/MS provides good selectivity and sensitivity for the detection of the oxidation product of 6B- 132 hydroxycortisol within the urine matrix. Results from the oxidation-based assay were compared to results obtained by a conventional GC-MS method based on formation of the methoxime-trimethylsilyl ether derivative of 6B-hydroxycortisol. The determination of 6B-hydroxycortisol in human urine was comparable by the two methods. The results of this research project indicate that low levels of dexamethasone and 6B—hydroxycortisol can be determined accurately using oxidation to convert the steroids to electrophilic analogs suitable for analysis by ECNI-MS. This methodology could be applied to other corticosteroid drugs and steroids that possess the 1,4-diene-3-one-11—oxy moiety. Also, steroids such as cholesterol and dehydroepiandrosterone, that possess a 5-ene-3-ol moiety, can be oxidatively converted to highly electrophilic 4-ene-3,6-dione analogs for analysis by ECNI-MS. A few of the oxidized plasma and urine matrix components produced large responses under ECNI-MS conditions. Identification of these components might lead to the development of sensitive assays for these compounds. In addition to oxidation, there are many selective organic reactions that could be utilized by the analytical chemist. For example, incorporation of the electrophilic 1,4- diene-3-one moiety into 3-hydroxy bile acids has been accomplished in a one-step synthesis using iodoxybenzene and benzeneselenic anhydride (1). This reaction could possibly be used to selectively enhance the electron capture response of 3-hydroxy steroids. 1. REFERENCES T. Iida, T. Shinohara, J. Goto, T. Nambara, and F. C. Chang, J. Lipid Res., 29, 1097—1 101, 1988. 133 "Ililllllllllllllllilllli